content
stringlengths
1
15.9M
\section{Introduction} Magnetic reconnection is the general process by which free magnetic energy stored in a plasma can be converted into kinetic and thermal energy by breaking the frozen-in constraint which exists for a perfectly conducting plasma. Magnetic reconnection occurs in a variety of astrophysical plasmas, including the interstellar medium (ISM), galactic disks, and the solar atmosphere \citep{2009ARA&A..47..291Z}. It has been suggested as a cause of many transient phenomena, such as solar flares and X-ray jets \citep{2011ApJ...731L..18M}, magnetospheric substorms \citep{2007JGRA..11206215B,2007JGRA..11201209M} and solar $\gamma$-ray bursts \citep{2005JGRA..11011103E,2005ApJ...628L..77T}. Reconnection also occurs in laboratory plasmas \citep{2010PhPl...17j2106G}, and plays a key role in the self-organization of fusion plasmas \citep{2006PhRvL..96s5003P,2007PhRvL..98g5001G}. Magnetic reconnection has long been considered as a mechanism for creating transient phenomena in the corona, such as solar flares, coronal mass ejections, and more recently X-ray jets (see review by \citet{2011ApJ...731L..18M}). In addition, recent observations of the solar chromosphere, the cooler, weakly ionized plasma below the solar corona, have suggested that reconnection is occurring there also. In particular, the quiet chromosphere exhibits localized, transient outflows on a number of different length-scales. The largest class of such outflows are called ``chromospheric jets'' \citep{2007Sci...318.1591S}. These are surges of plasma, observed in H$\alpha$ and Ca-II, with typical lifetimes of 200-1000 s, lengths of 5 Mm, and velocities at their base of 10 km/s. Analysis of observations of chromospheric jets with Hinode have shown ``blobs" of plasma within the jet outflow \citep{2007Sci...318.1591S}. A smaller class of these outflows are called ``spicules" \citep{2000SoPh..196...79S}, and are mainly observed on the solar limb, though possible disk counterparts have recently been identified in blue-shifts of Ca I and H$\alpha$ emission \citep{2012ApJ...752..108S}. Spicules have lifetimes of 10-600 s, lengths of up to 1 Mm and velocities of 20-150 km/s. A unified model of chromospheric outflow generation has recently been suggested by \citet{2007Sci...318.1591S} and \citet{2011ApJ...731L..18M}. The basic structure of the theoretical model consists of the so-called ``anemone" structure: a bipole field emerging into and reconnecting with a unipolar field. By changing the size of the emerging bipole, this simple model has been presented as a way of explaining spicules and chromospheric jets, as well as solar X-ray jets. Other evidence that magnetic reconnection is a possible driver of chromospheric transient phenomena includes Alfv\'{e}nic flows in spicules and ``blobs" of plasma in chromospheric jets, both of which are by-products of magnetic reconnection. A naive estimate of the reconnection rate $M$ (the rate at which magnetic field reconnects and ejects plasma) for these chromospheric phenomena can be made by dividing the inflow rate of plasma into the reconnection site by the outflow rate. The simple ``anemone'' structure described above has a flux inflow reservoir which is of the same size as the outflow region, and when this reservoir of flux is depleted, the reconnection stops. Therefore a typical size $L$ is used for both inflow and outflow regions, and given an outflow speed $v_{out}$, and a lifetime $t_{life}$, the reconnection rate can be estimated as $M\approx\frac{L }{v_{out}t_{life}}$. Taking the range of lifetimes, lengths and velocities for chromospheric jets gives a minimum of $M\approx 0.5$. Doing the same for spicules gives a minimum of $M\approx 0.01$. Traditionally, magnetic reconnection has been considered within the single-fluid, fully ionized, magnetohydrodynamic (MHD) framework, which is applicable to collisional plasmas. For a highly conducting plasma, the diffusion due to electron-ion collisions is negligible and the magnetic fieldlines are frozen into the plasma. Reconnection occurs when the frozen-in constraint is broken on timescales much shorter than the classical diffusion time, by allowing fieldlines to reconnect through a narrow diffusion region. \citet{1957JGR....62..509P} and \citet{Sweet58} were the first to formulate magnetic reconnection as a local process by considering a current layer of width $\delta$ much smaller than its length $L$, and with non-zero resistivity ($\eta$) due to electron-ion collisions. The model assumes steady state, i.e., the rate of plasma flowing into the diffusion region is equal to the rate of plasma flowing out, and takes the length $L$ to be the characteristic system length scale. Using Ohm's law for a fully ionized, single fluid plasma \begin{equation} \mathbf{E}+\mathbf{v}\times\mathbf{B} = \eta \mathbf{j}, \end{equation} and using the plasma momentum equation, a simple scaling law for the reconnection rate can be derived: \begin{equation} M \equiv \frac{v_{in}}{v_{A}} \approx \sqrt{\frac{\eta}{\mu_{0}v_{A}L}} = \frac{1}{\sqrt{S}}. \end{equation} Here $S=\mu_{0}v_{A}L/\eta$ is the Lundquist number, and $v_{A}=B/\sqrt{\rho\mu_{0}}$ is a typical Alfv\'{e}n velocity, with $B$ evaluated upstream from the current sheet and $\rho$ evaluated in the current sheet. From this analysis, it can also be shown that the current sheet aspect ratio $\sigma \equiv \frac{\delta}{L}$ scales as \begin{equation} \sigma \propto \frac{1}{\sqrt{S}}. \end{equation} Using the one dimensional (1D) semi-implicit model for the quiet Sun of \citet{1981ApJS...45..635V} gives a mass density of $7.4\times10^{-8} ~ \textrm{kg}/\textrm{m}^3$ for the chromosphere at 1 Mm above the solar surface. Using a magnetic field strength of 50 G gives an Alfv\'{e}n speed of approximately 15 km/s. The temperature at 1 Mm in this 1D model is $\approx$ 6000 K and the electron density is $\approx 10^{17} ~ \textrm{m}^{-3}$, which gives the electron-ion collision frequency to be $10^7 ~ \textrm{s}^{-1}$, and the Spitzer resistivity to be 0.004 $\Omega$m. Assuming a typical length scale of 0.1 Mm gives a Lundquist number of $S\approx10^6$ and a Sweet-Parker reconnection rate of $M\approx10^{-3}$. This is slower than the $M \ge 0.01$ required if reconnection is to explain the observed lifetimes of both chromospheric jets and spicules. The problem is greater in the corona, where $S\approx 10^{9}$, and the Sweet-Parker model predicts reconnection rates much too slow relative to those implied from observations. This is the general problem of the Sweet-Parker model: the predicted reconnection rate is significantly slower than almost all atmospheric transient phenomena. \citet{Petschek64} modified the Sweet-Parker reconnection formulation by allowing plasma to be redirected by standing shock waves which are setup at the ends of the diffusion region. This allowed for a shorter diffusion region, with length $L'$, and the reconnection rate increased by $\sqrt{L/L'}$. The maximum reconnection rate was found to be $M = \frac{\pi}{8 \ln(S)}$. For a value of $S=10^6$, this gives 0.065: substantially faster than the Sweet-Parker prediction of 0.001, and within the range observed for chromospheric spicules, though still too small for chromospheric jets. While the Petschek model predicts faster reconnection rates than the Sweet-Parker prediction, numerical simulations can only reproduce the Petschek reconnection regime if the resistivity is localized near the X-point - so-called \textit{anomalous resistivity} due to turbulence in the current layer or ion-cyclotron wave effects \citep{2003ApJ...587..450U,2004PhPl...11.2199S} - or if ion-electron drift terms are retained in the Ohm's law \citep{Birn01}. More recently, \citet{Loureiro07} and \citet{Huang10} considered the secondary tearing instability of the Sweet-Parker current sheet in a fully ionized plasma as an alternative means for obtaining fast reconnection. When the current sheet aspect ratio ($\sigma \equiv \delta/L$) reached a critical value of $\sigma_{c} = 1/200$, the sheet became unstable to the secondary tearing instability. Here, thinner current sheets were created between the primary plasmoids, which were themselves then prone to breaking into secondary plasmoids. With Sweet-Parker scaling in the laminar regime, they found that secondary onset occurred for a critical Lundquist number $S_{c}$ of $S_{c} = 1/\sigma_{c}^{2}= 4\times10^{4}$. Above this value, the reconnection rate was independent of $S$: $M=1/\sqrt{S_{c}}$. For a chromospheric Lundquist number of $S=10^{6}$, the reconnection rate as a result of the plasmoid instability gives $M=1/\sqrt{S_{c}}=5\times10^{-3}$, 5 times the Sweet-Parker prediction. Thus, the secondary instability allows reconnection independent of the mechanism which breaks the frozen-in constraint, but still gives reconnection rates slower than is needed to explain the parameters associated with chromospheric reconnection events. While secondary tearing may be the cause of the plasma ``blobs'' seen in chromospheric jets, until now no simulations have been performed to see if chromospheric magnetic reconnection could yield these plasmoids. This paper studies magnetic reconnection in the partially ionized chromosphere, focusing on both fast laminar reconnection and plasmoid formation. The plasma-$\beta$, the ratio of plasma pressure to magnetic pressure, can be as high as $10^{2}$ and as low as $10^{-4}$ in this region of the solar atmosphere \citep{2001SoPh..203...71G}. The average mass density falls over 4 orders of magnitude in a few Mm \citep{1981ApJS...45..635V} from the photosphere to the base of the corona. The ionization fraction, defined as the percentage of the plasma which is ionized, ranges from 0.1\% to 50\%, as the neutral density falls off faster than the ionized component density. In the bulk of the chromosphere the average collision time between neutral atoms and ions is of the order of ms \citep{1981ApJS...45..635V}. This is much less than a typical chromospheric time scale of 7~s, based on a typical Alfv\'{e}n velocity of 15~km/s and a typical length scale of 100~km. Hence the ions and neutrals are often treated as a single fluid. However, when magnetic diffusion length scales become as small as the neutral-ion collision mean free path, e.g., at magnetic reconnection sites, the decoupling of neutrals and ions cannot necessarily be neglected, and chromospheric magnetic reconnection should be studied in a multi-fluid framework. Partial ionization affects magnetic reconnection in several ways. The most obvious is the effect on the resistivity. Ion-neutral collisions introduce a Pedersen resistivity (or equivalently, an ambipolar diffusion) in the single-fluid MHD formulation, which acts perpendicular to the magnetic field \citep{1965RvPP....1..205B}. \citet{1994ApJ...427L..91B,1995ApJ...448..734B} showed that this ambipolar dissipation can create thin current structures in weakly ionized plasmas. The ionization level can also affect the reconnection rate through the Alfv\'{e}n speed: for coupled systems, the outflow Alfv\'{e}n speed depends on the total plasma density, but for decoupled systems it depends only on the ionized component density. Previously, \citet{1999ApJ...511..193V}, \citet{2003ApJ...583..229H}, and \citet{2004ApJ...603..180L} considered magnetic reconnection in the weakly ionized ISM, using a 1D analytic approach (note that the 1D paradigm assumes that outflow is negligible). In the reconnection inflow, if the ions that are pulled in by the reconnecting magnetic field are decoupled from the neutrals, an excess of ions can build up in the reconnection region. Recombination can then act as a sink for the ions, also decoupling them from the field. Given a sufficiently large recombination sink, the reconnection rate can become independent of the resistivity. This prediction has not been studied in higher dimensional magnetic reconnection configurations. Below, we examine this and other aspects of magnetic reconnection in a weakly ionized plasma within a two-dimensional (2D) numerical model. \citet{2008A&A...486..569S} and \citet{2009ApJ...691L..45S} performed two-fluid (ion+neutral) simulations of the coalescence of magnetic structures in partially ionized plasmas. They found that the reconnection rate for a fixed resistivity decreased as the plasma became less ionized, suggesting that jets associated with fast reconnection must occur in the upper chromosphere, where the ionization level is higher. However, in their investigations, the collision frequency and ionization/recombination rates did not depend self-consistently on the local plasma parameters (temperature and density), and therefore did not vary with space and time during the magnetic reconnection, a phenomena which is vital to test the predictions of \citet{1999ApJ...511..193V}, \citet{2003ApJ...583..229H}, and \citet{2004ApJ...603..180L}. In addition, their choice of resistivity was approximately 10 $\Omega$m, which is 4 orders of magnitude larger than a typical value in the chromosphere, and their Lundquist number was subsequently a relatively low value of approximately 10. In this paper, the first self consistent 2D simulations of chromospheric magnetic reconnection in a weakly ionized reacting plasma are performed. The numerical model used to simulate this reconnection is presented in \S2. The results are presented in \S 3, with particular focus on the scaling of the reconnection rate with resistivity, and the onset of the plasmoid instability. In \S 4 these results are used to evaluate the likelihood that magnetic reconnection can explain transient phenomena observed in the chromosphere such as spicules and jets. \section{Numerical Method} \subsection{Multi-Fluid Partially Ionized Plasma Model} \label{sec:HiFi_PN} A partially ionized reacting multi-fluid hydrogen plasma model is used to simulate reconnection in the solar chromosphere. For a detailed description and derivation of the model the reader is directed to \citet{Meier11} and \citet{Meier12b}. The model is implemented in the implicit, adaptive high order finite (spectral) element code framework, HiFi \citep{2008PhDT.........1L}. The full model consists of three fluids, ion (\textit{i}), electron (\textit{e}), and neutral (\textit{n}). The fluids can undergo recombination, ionization and charge exchange interactions, with $\Gamma_\alpha^r$ denoting the reaction rate for interaction $r$ affecting fluid $\alpha$. The recombination reaction rate for ions (the rate of change of ion number density due to recombination), $\Gamma_{i}^{rec}$, is defined as \begin{equation} \Gamma_{i}^{rec} \equiv -n_{i}\nu^{rec}, \end{equation} where the recombination frequency \begin{equation} \nu^{rec} = n_{e}\frac{1}{\sqrt{T_{e}^{*}}} 2.6\times10^{-19} \textrm{m}^{3}\textrm{s}^{-1} \label{eqn:recomb} \end{equation} is obtained from \citet{2003poai.book.....S} and $T_{e}^{*}$ is the electron temperature $T_e$ specified in eV. The ionization reaction rate for neutrals, $\Gamma_{n}^{ion}$, is defined as \begin{equation} \Gamma_{n}^{ion} \equiv -n_{n}\nu^{ion}, \label{eqn:ioniz} \end{equation} where the ionization frequency \begin{equation} \nu^{ion} = n_{e}A \frac{1}{X+\phi_{ion}/T_{e}^{*}}\left(\frac{\phi_{ion}}{T_{e}^{*}}\right)^{K}e^{-\phi_{ion}/T_{e}^{*}} \textrm{m}^{3}\textrm{s}^{-1} \end{equation} is given by the practical fit from \citet{1997ADNDT..65....1V}, using the values $A=2.91\times10^{-14}$, $K=0.39$, $X=0.232$, and the Hydrogen ionization potential $\phi_{ion}=13.6 \textrm{eV}$. Note that $\Gamma_{i}^{ion} = -\Gamma_{n}^{ion}$, and $\Gamma_{n}^{rec} = -\Gamma_{i}^{rec}$. The charge exchange reaction rate, $\Gamma^{cx}$ is defined as \begin{equation} \Gamma^{cx} \equiv \sigma_{cx}(V_{cx})n_{i}n_{n}V_{cx} \end{equation} where \begin{equation} V_{cx}\equiv \sqrt{\frac{4}{\pi}v_{Ti}^{2}+\frac{4}{\pi}v_{Tn}^{2}+v_{in}^{2}} \end{equation} is the representative speed of the interaction and $v_{in}^{2} \equiv {|\mb{v}_{i}-\mb{v}_{n}|}^2$ with $\mb{v}_{\alpha}$ denoting the velocity of species $\alpha$. The thermal speed of species $\alpha$ is given by $v_{T\alpha} = \sqrt{\frac{2k_{B}T_{\alpha}}{m_{\alpha}}}$, where $T_{\alpha}$ is the temperature, $m_{\alpha}$ is the corresponding particle's mass, and $k_{B}$ is Boltzmann's constant. Functional forms for the charge exchange cross-section $\sigma_{cx}(V_{cx})$ can be found in \citet{Meier11}. The multi-fluid model can be expressed by taking the neutral continuity equation, momentum equation, and energy (or pressure) equation to obtain a set of equations for the neutral fluid, and combining the electron and ion versions of these equations to obtain a set of equations for the ``ionized'' fluid. These equations are supplemented with Faraday's Law, and the required transport closure equations. Assuming charge neutrality in a hydrogen plasma, the electron and ion number densities are set equal to each other ($n_{i}=n_{e}$). Also the ionized and neutral atom masses are set equal to the proton mass ($m_{i}=m_{n}=m_{p}$). The resulting system of partial differential equations (PDEs) is given below. \nind\textit{Continuity: } \\ Due to charge neutrality, only the ion and neutral continuity equations are required. \bea \frac{\partial n_{i}}{\partial t} + \nabla \cdot (n_{i} \mb{v}_i) & = & \Gamma_i^{ion} + \Gamma_i^{rec}, \\ \frac{\partial n_n}{\partial t} + \nabla \cdot (n_n \mb{v}_n) & = & \Gamma_n^{rec} + \Gamma_n^{ion}. \eea \nind\textit{Momentum: } \\ The electron and ion momentum equations are summed and terms of order $(m_e/m_p)^{1/2}$ and higher are neglected to give: \bea \label{eq:mom_i} \frac{\partial}{\partial t} (m_i n_{i} \mb{v}_i) + \nabla \cdot (m_i n_{i} \mb{v}_i \mb{v}_i + \mathbb{P}_i+ \mathbb{P}_e) & = & \mb{j} \times \mb{B} + \mb{R}_i^{in} + \Gamma_i^{ion} m_i \mb{v}_n - \Gamma_n^{rec} m_i \mb{v}_i \nonumber \\ & +& \Gamma^{cx} m_i (\mb{v}_n - \mb{v}_i) + \mb{R}_{in}^{cx} - \mb{R}_{ni}^{cx}. \eea \nind The neutral momentum equation is \bea \label{eq:mom_n} \frac{\partial}{\partial t} (m_i n_n \mb{v}_n) + \nabla \cdot (m_i n_n \mb{v}_n \mb{v}_n + \mathbb{P}_n) & = & - \mb{R}_i^{in} + \Gamma_n^{rec} m_i \mb{v}_i -\Gamma_i^{ion} m_i \mb{v}_n \nonumber \\ & + & \Gamma^{cx} m_i (\mb{v}_i - \mb{v}_n) - \mb{R}_{in}^{cx} + \mb{R}_{ni}^{cx}. \eea \nind Here the current density is \begin{equation} \mathbf{j} = en_{i}(\mathbf{v}_{i}-\mathbf{v}_{e})=\frac{\nabla\times\mb{B}}{\mu_{0}}. \end{equation} The momentum transfer $\mb{R}_\alpha^{\alpha\beta}$ is the transfer of momentum to species $\alpha$ due to identity-preserving collisions with species $\beta$: \begin{equation} \mb{R}_\alpha^{\alpha\beta} = m_{\alpha\beta}n_{\alpha}\nu_{\alpha\beta}(\mb{v}_{\beta}-\mb{v}_{\alpha}), \end{equation} where $m_{\alpha\beta} = \frac{m_{\alpha}m_{\beta}}{m_{\alpha}+m_{\beta}}$. The collision frequency $\nu_{\alpha\beta}$ is given by \begin{equation} \nu_{\alpha\beta} = n_{\beta}\Sigma_{\alpha\beta}\sqrt{\frac{8k_{B}T_{\alpha\beta}}{\pi m_{\alpha\beta}}}, \end{equation} with $T_{\alpha\beta} = \frac{T_{\alpha}+T_{\beta}}{2}$. The cross-section $\Sigma_{in} = \Sigma_{ni}$ is $1.41\times 10^{-19} ~ \textrm{m}^2$, and the cross-section $\Sigma_{en} = \Sigma_{ne}$ is $1\times 10^{-19} ~ \textrm{m}^2$, assuming solid sphere elastic collisions \citep{1983ApJ...264..485D}. The cross section for ion-electron collisions, $\Sigma_{ei}= \Sigma_{ie}$, is assumed to be $\pi r_{d}^{2}$ where $r_{d}$ is the distance of closest approach ($e^{2}/(4\pi\epsilon_{0} k_{B}T_{ei})$ with $\epsilon_{0}$ the permittivity of free space and $e$ the elementary charge). The momentum transfer from species $\beta$ to species $\alpha$ due to charge exchange is $\mb{R}_{\alpha\beta}^{cx}$. As found by \citet{1995JGR...10021595P} and detailed in \citet{Meier11} and \citet{Meier12b}, appropriate approximations for these terms are \begin{equation} \mb{R}_{in}^{cx} \approx -m_{i}\sigma_{cx}(V_{cx})n_{i} n_{n}\mb{v}_{in}v_{Tn}^{2}\left[4\left(\frac{4}{\pi}v_{Ti}^{2}+v_{in}^2 \right)+\frac{9}{4\pi}v_{Tn}^{2}\right]^{-1/2}, \end{equation} and \begin{equation} \mb{R}_{ni}^{cx} \approx m_{i}\sigma_{cx}(V_{cx})n_{i} n_{n}\mb{v}_{in}v_{Ti}^{2}\left[4\left(\frac{4}{\pi}v_{Tn}^{2}+v_{in}^2 \right)+\frac{9}{4\pi}v_{Ti}^{2}\right]^{-1/2}. \end{equation} The pressure tensor is $\mbb{P}_\alpha = P_\alpha\mbb{I} + \pi_\alpha$ where $P_\alpha$ is the scalar pressure and $\pi_\alpha$ is the viscous stress tensor, given by $\pi_{\alpha} = -\xi_{\alpha}[\nabla\mb{v}_{\alpha} + (\nabla\mb{v}_{\alpha})^{\top}]$ where $\xi_{\alpha}$ is the isotropic dynamic viscosity coefficient for the fluid $\alpha$. \nind\textit{Internal Energy:} \\ Again, combining the electron and ion energy equations together and neglecting terms of the order $(m_e/m_p)^{1/2}$ and higher gives: \bea \label{eq:energy_i} \frac{\partial}{\partial t}\left(\ve_i + \frac{P_e}{\gamma-1}\right) & +& \nabla\cdot\left(\ve_i \mb{v}_i + \frac{P_e\mb{v}_e}{\gamma-1} + \mb{v}_i\cdot\mathbb{P}_i + \mb{v}_e\cdot\mathbb{P}_e + \mb{h}_i + \mb{h}_e\right) = \mb{j}\cdot\mb{E} \nonumber \\ &+& \mb{v}_i\cdot\mb{R}_i^{in} + Q_i^{in} - \Gamma_n^{rec}\frac{1}{2} m_i v_i^2 - Q_{n}^{rec} + \Gamma_i^{ion}(\frac{1}{2} m_i v_n^2 - \phi_{ion} ) + Q_i^{ion} \nonumber \\ & + & \Gamma^{cx}\frac{1}{2}m_i\left(v_{n}^{2} - v_{i}^{2}\right) + \mb{v}_n \cdot \mb{R}_{in}^{cx} - \mb{v}_i \cdot \mb{R}_{ni}^{cx} + Q_{in}^{cx} - Q_{ni}^{cx}. \eea The neutral energy equation is \bea \label{eq:energy_n} \frac{\partial \ve_n}{\partial t} &+& \nabla \cdot (\ve_n \mb{v}_n + \mb{v}_n \cdot \mathbb{P}_n + \mb{h}_n)\nonumber \\ & = & -\mb{v}_n \cdot \mb{R}_i^{in} + Q_n^{ni} - \Gamma_i^{ion}\frac{1}{2} m_i\mb{v}_n^2 - Q_i^{ion} + \Gamma_n^{rec}\frac{1}{2} m_i\mb{v}_i^2 + Q_n^{rec} \nonumber \\ &+& \Gamma^{cx} \frac{1}{2} m_i (\mb{v}_i^2 - \mb{v}_n^2) + \mb{v}_i \cdot \mb{R}_{ni}^{cx} - \mb{v}_n\cdot\mb{R}_{in}^{cx} + Q_{ni}^{cx} - Q_{in}^{cx}. \eea \nind Here $\ve_\alpha\equiv m_\alpha n_\alpha v_\alpha^2/2 + P_\alpha /(\gamma-1)$ is the internal energy density of fluid $\alpha$, and the term $\Gamma_i^{ion}\phi_{ion}$ represents optically thin radiative losses. The ratio of specific heats is denoted by $\gamma$. $Q_\alpha^{\alpha\beta}$ is the heating of species $\alpha$ due to interaction with species $\beta$, which is a combination of frictional heating and a thermal transfer between the two populations: $Q_\alpha^{\alpha\beta} = \mb{R}_\alpha^{\alpha\beta}\cdot(\mb{v}_{\beta} - \mb{v}_{\alpha}) + m_{\alpha\beta}n_{\alpha}\nu_{\alpha\beta}(T_{\beta}-T_{\alpha})$. The heat fluxes $\mb{h}_e$, $\mb{h}_{i}$ are calculated using the Braginskii closure for a magnetized plasma, and can be written as \begin{equation} \mb{h}_{\alpha} = \left[ \kappa_{\|,\alpha}\hat{\mb{b}}\hat{\mb{b}} + \kappa_{\bot,\alpha}(\mathbb{I}-\hat{\mb{b}}\hat{\mb{b}}) \right]\cdot\nabla k_{B} T_\alpha \end{equation} where $\kappa_{\|,\alpha}(T_\alpha)$ and $\kappa_{\bot,\alpha}(n_\alpha,T_\alpha,|\mb{B}|)$ account for the effects of thermal diffusion parallel to and perpendicular to the magnetic field direction ($\hat{\mb{b}}$) respectively, and whose functional forms can be found in \citet{1965RvPP....1..205B}. The neutral thermal conduction is isotropic $\mb{h}_{n} = -\kappa_{n}\nabla k_{B} T_{n}$. $Q_{\alpha}^{r}$ denotes thermal energy gain of species $\alpha$ due to a reaction \textit{r}, with $Q_{i}^{ion} = \Gamma_{i}^{ion}\frac{3}{2}k_{B}T_{n}$ and $Q_{n}^{rec} = \Gamma_{n}^{rec}\frac{3}{2}k_{B}T_{i}$. $Q_{\alpha\beta}^{cx}$ denotes heat flow from species $\beta$ to species $\alpha$ due to charge exchange \citep{Meier11,Meier12b}. The temperature of the neutrals is given by $T_{n} = P_{n}/(n_{n}k_{B})$; and the ion and electron temperatures are assumed to be equal such that $T_i = P_{i}/(n_{i}k_{B}) = P_{e}/(n_{e}k_{B}) = T_e$. \nind\textit{Ohm's Law:} \\ The generalized Ohm's law is given by the electron momentum equation: \bea \label{eq:mom_e} \frac{\partial}{\partial t} (m_e n_e \mb{v}_e) + \nabla \cdot (m_e n_e \mb{v}_e \mb{v}_e) & - & \Gamma_n^{ion} m_e \mb{v}_n + \Gamma_i^{rec} m_e \mb{v}_e \nonumber \\ & = &- e n_e (\mb{E} + \mb{v}_e \times \mb{B}) - \nabla\cdot\mathbb{P}_e + \mb{R}_{e}^{ei} + \mb{R}_{e}^{en}. \label{eqn:ohms1} \eea In the HiFi implementation all terms on the left hand side of Equation (\ref{eqn:ohms1}) are neglected, as in the chromosphere of the Sun, the electron inertial scale $c/\omega_{pe}$ is likely much smaller than magnetic diffusion length scales. However, the electron viscous stress tensor is preserved in order to represent the effects of microturbulence and 3D instabilities \citep{2011Natur.474..184C}, and also to damp the dispersive Whistler and kinetic Alfv\'{e}n waves at the shortest resolvable wavelengths. Note that the electron-neutral collision term $\mathbf{R}_{e}^{en}$ cannot be neglected. Using the identity $\mb{v}_{i}=\mb{v}_{e}+\mb{j}/en_{i}$, and the definition $\mathbf{w}\equiv\mathbf{v}_{i}-\mathbf{v}_{n}$, Eq.~(\ref{eqn:ohms1}) can then be written as \begin{equation} \mb{E} + (\mb{v}_i \times \mb{B}) = \eta\mathbf{j} + \frac{\mb{j}\times\mathbf{B}}{en_{i}} -\frac{1}{en_{i}}\nabla\cdot \mathbb{P}_{e} - \frac{m_{e}\nu_{en}}{e}\mathbf{w}, \label{eqn:ohms} \end{equation} where \begin{equation} \eta=\frac{m_{e}n_{e}(\nu_{ei}+\nu_{en})}{(e n_{e})^2} \end{equation} is the electron, or Spitzer, resistivity, which includes electron-ion and electron-neutral collisions. The system is closed by the use of Faraday's law $\frac{\partial B}{\partial t} = -\nabla\times\mb{E}$. It is worth noting how the model presented here differs from the partially ionized single-fluid model used in \citet{2006A&A...450..805L} and \citet{2007ApJ...666..541A}. The single-fluid approach implicitly assumes that the ions and neutrals are in ionization balance, which is determined by the average density and temperature, and the center of mass velocity used in the Ohm's law is an average over ions and neutrals. The Pedersen resistivity, present in this partially ionized single-fluid approach, is a consequence of this center of mass velocity. The multi-fluid model presented here follows the ions and neutrals separately, thus self-consistently including the interactions between ions and neutrals which are represented by the Pedersen resistivity in the single-fluid approach. A key advantage of the multi-fluid model used in this paper is that ions and electrons are allowed to be out of local thermodynamic equilibrium (LTE). This will be shown to be vital for the onset of fast reconnection in chromospheric plasmas. \subsection{Normalization} The equations are non-dimensionalized by dividing each variable ($C$) by its normalizing value ($C_{0}$). The set of equations requires a choice of three normalizing values. Normalizing values for the length ($L_{0}=1\times10^{5} ~ \textrm{m}$), number density ($n_{0}=3.3\times10^{16} ~ \textrm{m}^{-3}$), and magnetic field ($B_{0}=1\times10^{-3} ~ \textrm{T} $) are chosen. From these values the normalizing values for the velocity ($v_{0}=B_{0}/\sqrt{\mu_{0}m_{p}n_{0}}=1.20\times10^{5} ~ \textrm{m}/\textrm{s}^{-1}$), time ($t_{0}=L_{0}/v_{0} =0.83$ s), temperature ($T_{0}=m_{p}B_{0}^{2}/k_{B}\mu_{0}m_{p}n_{0} = 1.75\times10^{6}~\textrm{K}$), pressure ($P_{0}=B_{0}^2/\mu_{0}=0.80 ~ \textrm{Pa}$) and resistivity ($\eta_{0} = \mu_{0}L_{0}v_{0} = 1.5\times10^{4} ~ \Omega$m), can be derived. \subsection{Initial conditions and simplified equations} The simulation domain extends from -36$L_{0}$ to 36$L_{0}$ in the $x$ direction and -6$L_{0}$ to 6$L_{0}$ in the $y$ direction, with a periodic boundary condition in the $x$-direction and perfectly-conducting boundary conditions in the $y$-direction. The size of the domain has been chosen so that the boundaries and the particular boundary conditions do not affect any properties of the reconnection region centered and localized near $x=0, ~ y=0$. The initial neutral fluid number density is 200$n_{0}$ ($6.6\times 10^{18} ~ \textrm{m}^{-3}$), and the ion fluid number density is $n_{0}$ ($3.3 \times 10^{16} ~ \textrm{m}^{-3}$). This gives a total (ion+neutral) mass density of $1.11\times10^{-8} ~ \textrm{kg}.\textrm{m}^{-3}$, and an initial ionization level ($\psi_{i}\equiv n_{i}/(n_{i}+n_{n}$)) of 0.5 \%. The initial electron, ion and neutral temperatures are set to 0.005$T_{0}$ ($8750 \textrm{K})$. These initial conditions are consistent with lower to middle chromospheric conditions, based on 1D semi-empirical models of the quiet Sun \citep{1981ApJS...45..635V}.\\ In this plasma parameter regime, ion-neutral identity preserving collisions and charge exchange (CX) interactions are equally important and have a very similar effect of collisionally coupling neutral and ion fluids with a neutral-ion collision mean free path of $L_{ni}=v_{T,n}/\nu_{n,i}=140 ~ \textrm{m}$. However, the detailed CX physics is substantially more complicated and so for simplicity of interpretation CX interactions are neglected in this initial study and will be considered in future work. Charge exchange terms (terms with the superscript $cx$) in Equations (\ref{eq:mom_i})-(\ref{eq:energy_n}) are thus dropped. The ion inertial scale for these plasma parameters is $c/\omega_{pi}\approx 1 m$, which is much smaller than any scale of interest. Consequently, we neglect the Hall ($\mb{j}\times\mb{B}$) and the electron pressure tensor ($\nabla\cdot\mbb{P}_{e})$ terms in Equation (\ref{eqn:ohms}), the electron viscous stress tensor ($\nabla\cdot\pi_{e})$ in Equation (\ref{eq:mom_i}) and Equation (\ref{eq:energy_i}), and set electron velocity $\mb{v}_e$ equal to ion velocity $\mb{v}_i$ in Equation (\ref{eq:energy_i}). Similarly, \citet{2011ApJ...739...72M} showed that in the geometry of a reconnection current sheet electron-neutral collisions are only important in calculating resistivity, and so the term $\frac{m_{e}\nu_{en}}{e}\mathbf{w}$ in Equation (\ref{eqn:ohms}) is also dropped. With these simplifications, we solve the following set of governing PDEs: \nind\textit{Continuity: } \\ \bea \frac{\partial n_i}{\partial t} + \nabla \cdot (n_i \mb{v}_i) & = & \Gamma_i^{ion} + \Gamma_i^{rec}, \\ \frac{\partial n_n}{\partial t} + \nabla \cdot (n_n \mb{v}_n) & = & \Gamma_n^{rec} + \Gamma_n^{ion}. \eea \nind\textit{Momentum: } \\ \bea \frac{\partial}{\partial t} (m_i n_i \mb{v}_i) + \nabla \cdot (m_i n_i \mb{v}_i \mb{v}_i + \mathbb{P}_i+ P_e) & = & \mb{j} \times \mb{B} + \mb{R}_i^{in} + \Gamma_i^{ion} m_i \mb{v}_n - \Gamma_n^{rec} m_i \mb{v}_i, \\ \frac{\partial}{\partial t} (m_i n_n \mb{v}_n) + \nabla \cdot (m_i n_n \mb{v}_n \mb{v}_n + \mathbb{P}_n) & = & - \mb{R}_i^{in} + \Gamma_n^{rec} m_i \mb{v}_i -\Gamma_i^{ion} m_i \mb{v}_n . \eea \nind\textit{Internal Energy:} \\ \bea \frac{\partial}{\partial t}\left(\ve_i + \frac{P_e}{\gamma-1}\right) & + & \nabla\cdot\left(\ve_i \mb{v}_i + \frac{\gamma P_e\mb{v}_i}{\gamma-1} + \mb{v}_i\cdot\mathbb{P}_i + \mb{h}_i + \mb{h}_e\right) \\ & = & \mb{j}\cdot\mb{E} + \mb{v}_i\cdot\mb{R}_i^{in} + Q_i^{in} - \Gamma_n^{rec}\frac{1}{2} m_i v_i^2 - Q_{n}^{rec} \\ & +& \Gamma_i^{ion}(\frac{1}{2} m_i v_n^2 - \phi_{ion} ) + Q_i^{ion}, \\ \frac{\partial \ve_n}{\partial t} &+& \nabla \cdot (\ve_n \mb{v}_n \mb{v}_n \cdot \mathbb{P}_n + \mb{h}_n) = -\mb{v}_n \cdot \mb{R}_i^{in} + Q_n^{ni} \nonumber \\ &-& \Gamma_i^{ion}\frac{1}{2} m_i\mb{v}_n^2 - Q_i^{ion} + \Gamma_n^{rec}\frac{1}{2} m_i\mb{v}_i^2 + Q_n^{rec}. \eea \nind{\textit{Ohm's Law:}\\ \begin{equation} \mb{E} + (\mb{v}_i \times \mb{B}) = \eta\mathbf{j}. \end{equation} Note that to investigate the scaling of the reconnection rate with the Lundquist number (S), the Spitzer resistivity $\eta$ is made a parameter of the simulations. The values used for $\eta$ are $[0.5, 1,2,4,8,20]\times 10^{-5} \eta_{0}$, and so $\eta$ lies in the range [0.08,3] $\Omega$m. The viscosity coefficients for neutrals and ions are set to $\xi_{i} = \xi_{n} = 10^{-3}\xi_{0}$ and the neutral thermal conduction coefficient is $\kappa_{n }=4\times10^{-3}\kappa_{0}$. The normalizing constants are $\xi_{0} = m_{p}n_{0}L_{0}v_{0}$ and $\kappa_{0}=m_{p}n_{0}L_{0}v_{0}^3/T_{0}$. Note that for these plasma parameters the isotropic neutral heat conduction is much faster than any of the anisotropic heat conduction tensor components for either ions or electrons, with collisional ion-neutral heat exchange $Q_{i}^{in} \gg \nabla\cdot(\mathbf{h}_i + \mathbf{h}_e)$. Thus, thermal diffusion for all species is dominated by neutral heat conduction and is primarily isotropic. A Harris current sheet is used for the initial magnetic configuration, and is given here in terms of the in-plane magnetic flux $A_{z}$: \begin{equation} A_{z} = - B_{0}\lambda_{\psi}\ln{\cosh{(y/\lambda_{\psi})}} \end{equation} where $\mathbf{B} = \nabla \times A_{z}\hat{\mathbf{e}}_{z}$ and $\lambda_{\psi}=0.5L_{0}$ is the initial width of the current sheet. To provide the outward force to maintain this current sheet, both the ionized pressure $P_{p} =P_{i}+P_{e}$ and the neutral pressure $P_{n}$ are increased in the sheet: \bea P_{p}(y) & = & P_{p} + \frac{1}{2}\frac{F}{\cosh^{2}{(y/\lambda_{\psi})}}, \\ P_{n}(y) & = & P_{n} + \frac{1}{2}\frac{1-F}{\cosh^2{(y/\lambda_{\psi})}}, \eea where $F=0.01P_{0}$ is chosen to maintain an approximate ionization balance of 0.5\% inside the current sheet. These perturbations ensure that the total (ion and neutral) pressure perturbation balances the Lorentz force of the magnetic field in the current sheet: \begin{equation} 0 = -\nabla (P_{n}+P_{p}) + \mathbf{j}\times\mathbf{B}. \end{equation} At the same time, a relative velocity between ions and neutrals is required, so that the frictional force $\mb{R}_{i}^{in}$ can couple the ionized and neutral fluids and keep the fluids individually in approximate force balance: \begin{eqnarray} 0 & = & -\nabla P_{p} + \mathbf{j}\times\mathbf{B} + \mb{R}_{i}^{in}, ~ \textrm{and} \\ 0 & = & -\nabla P_{n} - \mb{R}_i^{in}. \end{eqnarray} To give this balance, we choose an initial ion velocity \begin{equation} {v_{i}}_{y}(y) = \frac{(F-1)}{n_i n_n\nu_{in}}\frac{\tanh{(y/\lambda_{\psi})}} {\lambda_{\psi}\cosh^{2}{(y/\lambda_{\psi})}}\frac{n_{0}^2v_{0}^2}{P_{0}}. \end{equation} To this initial steady state a small, local perturbation of the flux, $A_{z}^{1}$ is added to initiate the primary tearing instability and start the magnetic reconnection: \begin{equation} A_{z}^{1}= - \epsilon e^{-{\left(\frac{x}{4\lambda_{\psi}}\right)}^{2}}e^{-{\left(\frac{y}{\lambda_{\psi}}\right)}^{2}} \end{equation} where $\epsilon = 0.01B_{0}L_{0}$. The initial state is shown in Figure \ref{fig:IC1}. Only a subset of the domain is shown, and the $y$ axis is stretched by a factor of 40, so that thin structures can be clearly seen in the plots. Using the symmetry of the initial conditions, only the top right quadrant of the domain is simulated, with the use of appropriate boundary conditions. \section{Results} \subsection{Decoupling of inflow during magnetic reconnection} As mentioned in the previous section, $\eta$ is made a parameter of the simulations, and is not dependent on the local plasma parameters, so that a general scaling law of reconnection rate with resistivity can be derived. Six simulations are performed which have values of the resistivity of $[0.5, 1,2,4,8,20]\times 10^{-5} \eta_{0}$. Firstly, the generic processes that are evident in the simulations are highlighted by focusing on one particular value of $\eta=0.5\times10^{-5}\eta_{0}$. Later, a relationship between reconnection rate and $\eta$ is determined for the range of $\eta$ in these simulations. Figure \ref{fig:reconnection} shows the early evolution of the reconnection region resulting from the initial perturbation. The Harris current sheet undergoes the tearing instability, magnetic field reconnects at the X-point and flux is ejected. By $t=537.5 t_{0}$ a Sweet-Parker-like reconnection region has been formed. At this stage, ions are pulled in by the magnetic field and drag the neutrals with them. Figure \ref{fig:decoupling} shows the ion and neutral flow fields, on a smaller subdomain of the simulation, centered on the reconnection region. Panel \ref{fig:decoupling}a) shows the vertical velocities at time $t=537.5 t_{0}$, with ions on the top left quadrant and neutrals on the top right quadrant. Panel \ref{fig:decoupling}b) shows the magnitude of the horizontal velocity, again with ions on the top left quadrant and neutrals on the top right quadrant. The bottom two quadrants of panels a) and b) show the flow vectors for ions, on the left, and for neutrals, on the right, both scaled to the same magnitude. The vertical velocities in panel \ref{fig:decoupling}a) show that the ions and neutrals are decoupled, with the ions flowing faster into the reconnection region than the neutrals. The difference between neutral and ion velocity is approximately 90\% of the ion velocity. The horizontal velocities in panel \ref{fig:decoupling}b) show that the ion and neutral outflows (the strong horizontal velocity at x=+/-2.5$L_{0}$) are coupled. The difference between the ion and neutral outflow is negligible compared to the actual flow. The Alfv\'{e}n speed can be estimated using the upstream magnetic field of $B_{up}=0.6B_{0}$, taken at the point in the current sheet where the current amplitude reaches half its maximum value, and the total number density at the center of the sheet of 280$n_{0}$ to give $v_{A} = 0.035v_{0}$. The coupled outflow of ions and neutrals, which has a maximum of $0.015v_{0}$, is thus approximately half the Alfv\'{e}n speed (based on total number density, as the outflow is coupled) at this time. Later in time the outflow increases to the Alfv\'{e}n speed. This feature of the $\eta=5\times10^{-6}\eta_{0}$ simulation is common to all six simulations. The ions and neutrals inflows are decoupled, as ions are pulled in by reconnecting magnetic field and the neutrals are dragged in via collisions. The timescale of inflow is fast enough that the collisions cannot keep the neutrals completely coupled to the ions and an excess of ions builds up in the reconnection region, creating an ionization imbalance. This is the situation considered for astrophysical plasmas by both \citet{1999ApJ...511..193V}, who treated weakly ionized reconnection with an analytic approach, and \citet{2003ApJ...583..229H}, who calculated analytic and numerical solutions of 1D steady state models of weakly ionized reconnection. \subsection{Reconnection rate scaling with resistivity} As described in the previous section, the plasma in the reconnection region is out of ionization balance as the neutrals are largely left behind by the ions. Figure \ref{fig:steady_state} shows the four components contributing to $\frac{\partial n_i}{\partial t}$ in the ion continuity equation. The top left quadrant shows the loss due to recombination, the bottom left quadrant shows the loss due to outflow (horizontal gradient in horizontal momentum of ions), the top right quadrant shows the gain due to inflow (vertical gradient in vertical momentum), and the bottom right shows the gain due to ionization. The color scheme is based on a log-scale, and shows that within the reconnection region gains due to ionization are negligible relative to the other three terms. Looking at the largest values for the remaining three terms, the losses due to recombination and outflow are comparable, and add up to equal the gain due to inflow. This shows that the reconnection region is close to a steady state, with inflow of ions balanced by comparable contributions from recombination and outflow. Recall that the 1D models of \citet{1999ApJ...511..193V} and \citet{2003ApJ...583..229H} assumed that recombination was fast enough to dominate over the outflow, and thus the horizontal direction (along the current sheet) was ignorable. Instead, Figure \ref{fig:steady_state} shows that for the self-consistently created reconnection region in this parameter range the recombination and outflow are comparable and so 2D effects cannot be neglected. The nature of the steady state balance in these reconnection simulations has important consequences for the scaling of the reconnection rate. In the standard Sweet-Parker reconnection scenario, inflow of ions into the reconnection region is assumed to be balanced by outflow of ions. This, along with other assumptions, gives the standard scaling of the normalized inflow rate (reconnection rate) of $ M\propto \sqrt{\eta} \propto 1/\sqrt{S}$. For the weakly ionized plasma in these simulations, the situation is very different due to the ionization imbalance. The standard steady-state argument of the Sweet-Parker model can be modified to include the reacting multi-fluid equations, in order to derive a scaling relationship for the reconnection rate. Figure \ref{fig:SP_PIP} shows a cartoon of the reconnection region in these simulations. The reconnection region, inside which the frozen-in constraint is broken by resistivity $\eta$, has width $\delta$ and length $L$. There is an external ion density of $n_{i,ext}$ and a current sheet ion density of $n_{i,CS}$. The inflow of ions into the reconnection region is $v_{in}$, the outflow of ions is $v_{out}$, and the upstream magnetic field is $B_{up}$. Assuming that the system is in steady state, i.e., that $\frac{\partial n_{i}}{\partial t}=0$, and integrating around the reconnection region gives \begin{equation} n_{i,ext}v_{in}L = n_{i,CS}(\delta v_{out} + \delta L \nu^{rec} - \delta L \nu^{ion}), \end{equation} where $\nu^{rec}$ and $\nu^{ion}$ are the recombination and ionization frequencies, defined in Equations (\ref{eqn:recomb}) and (\ref{eqn:ioniz}). Defining $\nu_{out} \equiv v_{out}/L$, and equating Ohm's law evaluated inside (${E}=\eta{j}$) and outside (${E}=v_{in}B_{up}$) of the reconnection region, \begin{equation} v_{in}B_{up} = \eta j \approx \frac{\eta B_{up}}{\delta \mu_{0}}, \end{equation} to eliminate $\delta$, this steady state equation can be rewritten as \begin{equation} v_{in} \approx \sqrt{ \frac{\eta}{\mu_{0}}\frac{n_{i,CS}}{n_{i,ext}} (\nu_{out} + \nu^{rec} - \nu^{ion})}. \label{eqn:rates} \end{equation} For a plasma in ionization balance, $\nu^{rec} = \nu^{ion}$, and $n_{i,ext} \approx n_{i,CS}$. Using these relationships and the total momentum equation to derive $v_{out} = v_{A}$ (where $v_{A} = \frac{B_{up}}{\sqrt{\mu_{0}\rho_{total}}}$), recovers the standard Sweet Parker scaling law \begin{equation} M\equiv \frac{v_{in}}{v_{A}} \approx \sqrt{\frac{\eta}{v_{A}L\mu_{0}}} = \sqrt{\frac{1}{S}}, \end{equation} where $B_{up}$ is the magnetic field upstream of the current sheet, and $\rho_{total}$ is the total (ion + neutral) density in the reconnection region. In contrast, in these reacting two-fluid simulations, the system non-linearly and self-consistently forms a current layer where plasma is out of ionization balance, where the recombination is comparable to the outflow, and where ionization is negligible compared to both recombination and outflow. Hence the inflow rate can be approximated by \begin{equation} v_{in} \approx \sqrt{ \frac{\eta}{\mu_{0}}\frac{n_{i,CS}}{n_{i,ext}}(\nu_{out} + \nu^{rec}) }. \end{equation} Note that this equation does not by itself indicate how the reconnection rate will depend on $S$, as it is not clear how $\nu^{rec},n_{i,CS}$, and $n_{i,ext}$ depend on $S$ from the equations. We will therefore use our series of simulations performed over a range of $\eta$ to determine the reconnection rate dependence on $S$. For all six simulations the effective Lundquist number is defined by \begin{equation} S_{sim} \equiv \frac{v_{A,0}L_{sim}\mu_{0}}{\eta}, \end{equation} where $L_{sim}=2.75L_{0}$. Note that $L_{sim}$ is much smaller than the horizontal extent of the domain. We define $v_{A,0}$ as the Alfv\'{e}n velocity given a field strength of $B_{0}$ and number density $201n_{0}$, i.e., based on the initial background magnetic field and number density. Note that $S_{sim}$ varies over simulations due to $\eta$ only. The reconnection rate is defined by \begin{equation} M_{sim} \equiv \frac{\eta j_{max}}{v_{out}B_{up}}. \end{equation} Here, $j_{max}$ is the maximum value of the current density, located at $(x,y)=(x_{j},0)$, within the reconnection region, and $B_{up}$ is evaluated at $(x_{j},\delta_{sim})$ where $\delta_{sim}$ is the $y$ location on the line $x=x_{j}$ at which the current density reaches $j_{max}/2$, i.e., half-width at half-max of the current sheet. For each simulation, the value of $M_{sim}$ is taken at a time when the length of the current sheet, defined by the distance from the original X-point to the location of maximum outflow $v_{out}$ equals $L_{sim}$. This happens at a different time for each simulation, but in all cases prior to the onset of any secondary instabilities of the current sheet. This instantaneous value of $M_{sim}$ is plotted for each simulation against the effective Lundquist number $S_{sim}$ in Figure \ref{fig:rec_rate}. The dashed line shows the single-fluid Sweet-Parker predicted scaling $M \propto 1/\sqrt{S}$. Over the range of $S$ which these six simulations cover, the reconnection rate is only weakly dependent on the Lundquist number, with a slow decrease in $M_{sim}$ with increasing $S_{sim}$. The reconnection rate is much faster than the Sweet-Parker reconnection rate, as a direct result of the decoupling of ions from neutrals and the enhanced recombination in the reconnection region. As discussed in section \ref{secondary}, plasmoid formation increases the reconnection rate for $S_{sim} > S_{c}$. We also note that each of these simulations have been performed with the same viscosity coefficient, and it is known that the reconnection rate dependency on resistivity weakens when viscous effects in the reconnection layer become important \citep{1984PhFl...27..137P}. Figure \ref{fig:width} shows the scaling of the reconnection region aspect ratio $\sigma_{sim}\equiv (\delta_{sim}/L_{sim})$ with $S_{sim}$ for the six simulations. Also shown is the Sweet-Parker scaling of $1/\sqrt{S}$. The simulations do not exhibit the $\sigma \propto 1/\sqrt{S}$ scaling but show an approximate $1/S$ scaling (the power law fit through the simulation points has an exponent of $-1.1 \pm 0.17$). This $1/S$ scaling is consistent with the result of Figure \ref{fig:rec_rate}, that $M_{sim}$ is approximately independent of $S_{sim}$. This is shown as follows: Assume that $v_{out}$ does not substantially vary over different simulations, and note that $S_{sim}$ only changes over simulations due to $\eta$. Then using the definition of $M_{sim}$ and $j_{max} \approx B_{up}/\delta_{sim}$ gives $M_{sim} \propto \eta j_{max}/B_{up} \approx \eta/\delta_{sim} \propto 1/S_{sim}\sigma_{sim} \approx \textrm{const.}$ \subsection{The secondary (plasmoid) instability} \label{secondary} As discussed above, at a critical aspect ratio of $\sigma_{c}=1/200$, a resistive current sheet can become unstable to a secondary tearing instability known as the \textit{plasmoid instability}. Of the six two-fluid simulations in this paper, three cases show evidence of secondary tearing. Figure \ref{fig:plasmoid} shows the onset of the plasmoid instability for the simulation with $\eta=0.5\times10^{-5}\eta_{0}$, with the two contours of magnetic flux at $A_{z}=-0.0681B_{0}L_{0}$ (white line) and $A_{z}=-0.0687B_{0}L_{0}$ (black line) following the evolution of two particular field-lines in time. The initial laminar current sheet breaks up into a number of plasmoids, with thinner current sheets between them. The current density and recombination rate within this fragmented reconnection region are shown in Fig.~\ref{fig:plasmoid} on a pseudo-color scale. During this phase of the reconnection the formation of a plasmoid chain redistributes the ionized plasma within the current sheet, which is otherwise approximately uniformly distributed throughout the reconnection region, into regions of higher and lower electron number density. Since $\Gamma_{i}^{rec}\propto n_e^2$, the recombination rate is increased within the plasmoids with respect to the sub-layers between them. This leads to the plasmoid magnetic flux collapsing on itself (i.e, disappearing as it would in vacuum) at a rate comparable to or faster than the plasmoids are exhausted out of the reconnection region. This collapse can affect the expected distribution of plasmoid sizes and the ionization fraction within the plasmoids relative to the background medium in the reconnection exhaust. \citet{Huang10} found that in a fully ionized plasma, the onset of the plasmoid instability occurred at a critical value of Lundquist number of $4\times10^{4}$ which, with the Sweet-Parker scaling of current sheet width with Lundquist number, corresponds to a critical current sheet aspect ratio of $1/200$, as shown by the dashed line in Figure \ref{fig:width}. The three simulations that undergo secondary tearing are shown as diamonds in Figure \ref{fig:width}. The intersection of a power law fit to the $\sigma_{sim}(S_{sim})$ data intersects the $\sigma_{sim}=1/200$ line at a value of $S_{sim}=10^{4}$. The simulation with $S_{sim}=10^4$ exhibits the plasmoid instability, as do the two simulations with higher $S_{sim}$ (shown as diamonds in the Figure). These two facts support the postulation of \citet{Loureiro07} that the criterion for the plasmoid instability is that the aspect ratio decreases below a critical value. For our simulations, as for \citet{Huang10}, this critical value is 1/200. To demonstrate the change in reconnection rate due to the plasmoid instability, $M_{sim}$ is evaluated again later in each of the three plasmoid-unstable simulations. It is difficult to measure $M_{sim}$ during the plasmoid instability, as the reconnection rate varies with time, depending on the plasmoid evolution. To give an idea of how the plasmoid instability is affecting the reconnection rate we plot in red in Figure \ref{fig:rec_rate} the range of $M_{sim}$ observed after plasmoid formation in each simulation until the current sheet width becomes spatially unresolved. It is apparent, particularly in higher $S_{sim}$ simulations, that the reconnection rate increases as the plasmoids develop. Thus, the reconnection rate in this regime is determined by both the fast recombination of ions, and the effect of the secondary tearing instability. \section{Discussion} In this paper magnetic reconnection in the solar chromosphere is simulated using a partially ionized reacting multi-fluid plasma model. The number densities and ionization levels are consistent with lower to middle chromospheric conditions. The dependence of reconnection rate on Lundquist number is investigated by setting the resistivity to be a parameter of the simulations. A simple 2D Harris current sheet configuration with a local perturbation to the in-plane flux is used. The system self-consistently and non-linearly creates a reconnection region which is out of ionization balance, due to the decoupling of ion and neutral inflows. The model is able to capture this physical effect as the two fluids are followed separately, and the ionization and recombination rates are self-consistently calculated based on local plasma parameters. In the reconnection region, recombination of excess ions is comparable to the outflow of ions, which leads to fast reconnection independent of the Lundquist number, and when normalized properly, the reconnection rate is approximately 0.1. It is worth noting that guide (out of plane) magnetic field, not included in these simulations, will have an effect on the scaling, as the flux associated with the guide field can inhibit fast reconnection \citep{2003ApJ...590..291H}. The onset of fast reconnection in weakly ionized astrophysical plasmas was predicted by \citet{1999ApJ...511..193V} and \citet{2003ApJ...583..229H}, assuming that recombination dominates outflow in the current sheet so the system could be treated with one dimensional analysis. In this regime \citet{2003ApJ...583..229H} found that fast reconnection occurred when \begin{equation} Z=\frac{\beta_0^{3/\gamma}}{10}\frac{t_{recomb}t_{\Omega}}{t_{AD}^2} < Z_{c} \end{equation} where $Z_{c}$ was either $10^{-4}$ for the numerical solution, or $10^{-2}$ for the theoretical solution, $\beta_0\equiv\frac{P_i+P_e}{B^2/\mu_0}$, $t_{rec}$ is the timescale of recombination, $t_{\Omega} \equiv L^2/\eta$, and $t_{AD} \equiv L^2/\eta_{p}$. In our multi-fluid simulations, the smallest resistivity is $\eta \approx 0.08 ~ \Omega \textrm{m}$, $\eta_{p} \approx 0.3 ~ \Omega \textrm{m}$, and $t_{rec} =1/\nu^{rec} \approx 30 ~ \textrm{s}$, and so $Z\approx4\times10^{-5}$, which lies in the `fast' reconnection regime of \citet{2003ApJ...583..229H}. While we have shown that while the prediction of fast reconnection by \citet{1999ApJ...511..193V} and \citet{2003ApJ...583..229H} holds in the chromosphere, in 2D reconnection the recombination is not the dominant mechanism for the removal of ions from the reconnection region, as outflows are equally important. The critical aspect ratio at which secondary tearing sets in, $\sigma_{c}=1/200$, is reached at a lower Lundquist number than has previously been seen in single-fluid simulations with $\beta \approx 1$ \citep{Huang10,Samtaney09}. This is because in this fast reconnection regime $\sigma \propto 1/{S}$, compared to Sweet-Parker reconnection where $\sigma \propto 1/\sqrt{S}$. Note that \citet{2012PhPl...19g2902N} found that for $\beta = 50$, the onset occurs at a Lundquist number of 2000-3000, and an aspect ratio of 1/60. The plasmoids which form due to secondary tearing in these simulations are losing ions due to recombination, and so their evolution is potentially very different from those seen in fully ionized simulations. The further evolution of the plasmoid beyond the initial formation and collapse seen here is left to a follow-up investigation. It has been conjectured that current sheets are ubiquitous in the chromosphere \citep{2012ApJ...751...75G}. It has also been established that spicules and chromospheric jets are ubiquitous solar phenomena \citep{2000SoPh..196...79S}. Magnetic reconnection in chromospheric current sheets is therefore a promising mechanism to link these two ubiquitous phenomena, and would explain the formation of spicules and chromospheric jets. These simulations show that the chromosphere can exhibit fast reconnection with rates that are comparable to the estimated reconnection rates of observed outflows, as well as the creation of plasmoids, or ``blobs" of plasma, due to the secondary tearing mode. The reconnection outflow velocities we find in these simulations are 5 km/s, which is close to the speeds observed for chromospheric jets, but is small compared to the speeds of 20-150 km/s observed for spicules. Using the scaling argument that $v_{out} \propto B_{up} /\sqrt{n_i+n_n}$, we expect that by increasing the magnetic field from 10G to 50G, and by moving higher up in the chromosphere where the neutral density is an order of magnitude lower (and the ionization level increases from 0.5\% to 10\%) spicule-like outflow velocities of up to 80 km/s can realistically be achieved. The current sheets formed in these simulations of chromospheric reconnection are out of ionization balance to such a degree that short recombination times of 30 s are created. These times are much smaller than the estimated 3 minute recombination time of an acoustic shock heated chromosphere \citep{2002ApJ...572..626C}, highlighting magnetic reconnection as the prime transient phenomena in the chromosphere driving non-LTE recombination. The major advantage of the multi-fluid approach over single-fluid models is that it can self consistently produce non-LTE high ion density structures. This paper has shown that such ion density structures form in magnetic reconnection, and that the resulting fast recombination affects the reconnection physics. Hence multi-fluid simulations such as these are vital to understanding transient phenomena in the chromosphere. Finally, it should be noted that \citet{2004PhPl...11.2199S} and others have argued that the addition of the ion inertial and/or finite Larmor radius effects to single-fluid resistive MHD is essential to obtain the resistivity-independent ``universal'' fast reconnection rate of $M\approx 0.1$. The simulation results presented here yield this same resistivity-independent reconnection rate without any of the ion-electron decoupling effects; relying instead on the combination of enhanced recombination rate with generation of the secondary plasmoids in the reconnection region. \begin{acknowledgments} \nind{Acknowledgements:} This work has been supported by the NASA Living With a Star \& Solar and Heliospheric Physics programs, the ONR 6.1 Program, the U.S. DOE Experimental Plasma Research program, by LLNL under Contract DE-AC52-07NA27344, and by the NRL-\textit{Hinode} analysis program. The simulations were performed under a grant of computer time from the DoD HPC program. \end{acknowledgments}
\section{Introduction} \label{sec:introduction} The local density approximation (LDA) is the simplest functional in density functional theory (DFT) and has been around since the advent of DFT~\cite{HohenbergKohn1964, KohnSham1965b}. Although the LDA has only a local dependence on the density, it has been tremendously successful in describing the ground-state properties of solids, surfaces and large molecules. Its shortcomings have been obvious from the beginning and an enormous effort has gone into the search for better approximations. This task has proved to be exceedingly difficult. One should, however, not be surprised or disappointed. Density functional theory provides a way to reduce the full, interacting many-body problem to a simple non-interacting one. Therefore, an accurate density functional for the total energy would provide a surprisingly simple way to solve the complicated many-body problem --- at least as far as static properties are concerned. Nevertheless, in the past decades considerable progress has been made. With the advent of the generalized gradient approximations (GGAs)~\cite{Becke1988, Perdew1991}, bond lengths and atomization energies were greatly improved as compared to those of the LDA. Unfortunately, the exchange-correlation (xc) potentials of the GGAs have several undesirable features. In particular, they decay too fast outside finite systems~\cite{LeeZhou1991, EngelChevaryMacdonald1992}, unlike a correct $-1/r$ decay. Consequently, neither the LDA nor the GGAs produce proper Rydberg levels~\cite{LeeuwenBaerends1994} and ionization potentials are too low. An important class of extensions to the GGAs came from the observation that exchange energies are much larger than correlation energies. This suggests that exchange should be treated exactly, while using a GGA only for the correlation energy. Full inclusion of ``exact exchange'' does however not work well in practice, since there is a large cancellation of errors between the LDA or the GGA versions of the exchange and correlation energies. On the other hand, using only a portion of exact exchange combined with a GGA does yield quite accurate bond energies in molecules~\cite{Becke1993}. An important advantage of ``exact exchange'' is that it provides some necessary improvements of the xc potential. For instance, the inclusion of ``exact exchange'' gives a stepped structure in the xc potential~\cite{LeeuwenGritsenkoBaerends1995, LeeuwenGritsenkoBaerends1996} thus improving the description of the atomic shells as well as enabling a neutral dissociation of heteronuclear molecules~\cite{PerdewParrLevyBalduz1982, GritsenkoBaerends1996, BaerendsGritsenko1997}. It also gives an xc-potential with the proper asymptotic ($-1/r$) behavior which gives rise to Rydberg levels and a highest occupied KS eigenvalue in better agreement with the negative of the ionization potential. Usually, however, only a fraction of full exchange is incorporated in the so called hybrid functionals meaning that the desirable features mentioned above are only partially obtained. It seems that only functionals with a massive amount of fitting parameters are able to handle 100\% ``exact exchange''~\cite{Gill2001, ZhaoTruhlar2008}, though they will always suffer from weak singularities in the response functions of metals. We mention here that inclusion of ``exact exchange'' is not mandatory in order to have good properties of the xc potential. The proper step structure as well as the correct asymptotics away from finite systems can also be obtained by modeling the potential directly~\cite{LeeuwenBaerends1994, GritsenkoLeeuwenLenthe1995}. Such model potentials can indeed provide good response properties~\cite{GritsenkoSchipperBaerends1999, SchipperGritsenkoGisbergen2000} but they cannot easily be written as the functional derivative of some accurate functional for the xc energy. As a result, their implementations have been limited. Further improvements to the energy functional are also sought by adding the kinetic-energy density to the functional arguments~\cite{GhoshParr1986, TaoPerdewStaroverov2003} and including parts from many-body perturbation theory~\cite{GorlingLevy1994, Leeuwen1996, GattiOlevanoReining2007, RomanielloSangalliBerger2009, HellgrenBarth2010}. In this paper we would like to consider an alternative approach by looking directly at electron correlations in real space. In a correlated system we can consider the conditional probability of finding an electron at some point in space, when the position of another reference electron is given. The difference between this function and the unconditional probability (which is simply the density) is defined to be the xc hole and the knowledge of this function~\footnote{To be precise we need to know xc hole for different strengths of the electron interaction.} is sufficient to calculate the xc-energy. The effect of exchange and correlation is to dig a hole in the density around each electron, so as to remove one electron in that region. We can say that the task of a good xc functional is to provide an accurate description of the xc hole. The LDA and GGA assume that this hole has a spherical shape and an extent given by the Wigner--Seitz radius ($4\pi r_s^3 n = 3$, $n$ being the electron density) at the reference electron. Unfortunately for the LDA, xc holes are definitely not spherical~\cite{PhD-Buijse1991, BuijseBaerends2002}. Although hybrid functionals lift this restriction, their blunt use of ``exact exchange'' actually worsens the description of the xc hole for stretched bonds compared to the LDA and GGA functionals. A more sophisticated class of functionals which aims to build a non-spherical model of the xc hole is the so-called weighted density approximations (WDA)~\cite{AlonsoGirifalco1978, GunnarssonJonsonLundqvist1979, SaddTeter1996, RushtonTozerClark2002}. These functionals avoid the spherical xc hole by digging a hole out of the real density rather than in the density at the position of the reference electron. A nice feature of the WDA is that the asymptotic behavior of its xc potential has a Coulombic asymptotic decay instead of an exponential behavior as in the LDA and the GGA. An important symmetry of the pair-density ($\Gamma(\mat{r}',\mat{r}) = \Gamma(\mat{r},\mat{r}')$) is, however, broken. This causes an additional factor $1/2$ in the asymptotic decay of the xc potential, so that it decays too fast (as $-1/(2r)$)~\cite{OssiciniBertoni1985}. Here, we will advocate an approach in a similar spirit as the weighted density approximation~\cite{Barth2004}. However, we will take care not to destroy the symmetry of the pair-density and therefore, the xc potential will automatically have the correct asymptotic $-1/r$ behavior. Furthermore, important information on the physics of the xc hole is provided by the dissociation of molecules. In particular, a proper localization of the xc hole around the reference electron for a dissociated molecule is a challenging task. The failure of current approximations to achieve this is reflected in their consistent inability to properly describe the breaking of chemical bonds. The most important aim of our new functional will therefore be a bold one: the new functional has to be able to describe molecular dissociation. The paper is outlined as follows. First (Sec.~\ref{sec:motivation}) we will give a more detailed discussion on the background to motivate the construction of the screened exchange (SX) functional. The actual construction of the SX functional is discussed in section~\ref{sec:SXmodel}. In section~\ref{sec:results} we show preliminary results and finally conclude in section~\ref{sec:conclusion}. \section{Motivation} \label{sec:motivation} \subsection{Symmetry and asymptotics} \label{sec:SymAndAsymp} We will start from an exact expression~\cite{Almbladh1972, LangrethPerdew1975, GunnarssonLundqvist1976} for the exchange-correlation energy \begin{align}\label{eq:xcEnergy} E_{\text{xc}} = \frac{1}{2} \iinteg{\mat{r}}{\mat{r}'}n(\mat{r})\frac{n(\mat{r}')\bigl(\bar{g}(\mat{r},\mat{r}') - 1\bigr)}{\abs{\mat{r}-\mat{r}'}}, \end{align} where $n(\mat{r})$ is the density. This expression gives the exact xc contribution to the interaction energy of the system, provided $\bar{g}$ is the exact pair-correlation function $g$ of the system, defined as \begin{align*} g(\mat{r},\mat{r}') \equiv \frac{\Gamma(\mat{r},\mat{r}')}{n(\mat{r})n(\mat{r}')}. \end{align*} Here the diagonal of the two-body reduced density matrix is defined as \begin{align*} \Gamma(\mat{r},\mat{r}') &\equiv \sum_{\sigma,\sigma'}\Gamma(\mat{r}\sigma,\mat{r}'\sigma') \notag \\ &\equiv \sum_{\sigma,\sigma'} \brakket{\Psi}{\crea{\psi}(\mat{r}\sigma)\crea{\psi}(\mat{r}'\sigma') \anni{\psi}(\mat{r}'\sigma')\anni{\psi}(\mat{r}\sigma)}{\Psi}, \end{align*} where $\crea{\psi}(\mat{r}\sigma)$ and $\anni{\psi}(\mat{r}\sigma)$ are the usual field operators. The xc energy, however, also contains a correlation contribution to the kinetic energy which is most conveniently included by integrating the interaction energy with respect to the strength $\lambda$ of the Coulomb interaction --- while keeping the density fixed at the fully interaction one ($\lambda=1$)~\cite{Almbladh1972, LangrethPerdew1975, GunnarssonLundqvist1976, Barth2004}, i.e. \begin{align*} \bar{g}(\mat{r},\mat{r}') \equiv \int_0^1\!\!\!\mathrm{d}\lambda\, g_{\lambda}(\mat{r},\mat{r}'). \end{align*} The physical picture of representing the xc energy in this manner is that an electron does not interact with the full density, but depending on its position, $\mat{r}$, it sees an effective density $n(\mat{r}')\bar{g}(\mat{r},\mat{r}')$ with $N - 1$ electrons. Therefore, the part \begin{align}\label{eq:xcHoleDef} \bar{\rho}_{\text{xc}}(\mat{r}'|\mat{r}) \equiv n(\mat{r}')\bigl(\bar{g}(\mat{r},\mat{r}') - 1\bigr) \end{align} in the xc energy~\eqref{eq:xcEnergy} exactly describes the density of minus one electron, a hole, which is reflected in the sum-rule of the xc hole \begin{align}\label{eq:xc holeSumRule} \integ{\mat{r}'} \bar{\rho}_{\text{xc}}(\mat{r}'|\mat{r}) = \integ{\mat{r}'} n(\mat{r}')\bigl(\bar{g}(\mat{r},\mat{r}') - 1\bigr) = -1. \end{align} The local density approximation (LDA) proceeds by using the xc hole from the homogeneous electron gas evaluated for the density at the position of the reference electron, so the pair-correlation function is approximated as $\bar{g}(\mat{r},\mat{r}') \approx \bar{g}_h\bigl(\abs{\mat{r}-\mat{r}'};n(\mat{r})\bigr)$. Furthermore, if the distance $\abs{\mat{r}-\mat{r}'}$ is large the pair-correlation function $\bar{g}$ only differs slightly from one, so it is quite reasonable to replace $n(\mat{r}')$ by $n(\mat{r})$ in the xc-energy~\eqref{eq:xcEnergy}. Combining both approximations, we obtain the expression for the xc energy of the LDA \begin{align}\label{eq:LDAenergy} E_{\text{xc}}^{\text{LDA}} &= \frac{1}{2} \iinteg{\mat{r}}{\mat{r}'}n(\mat{r}) \frac{n(\mat{r})\bigl\{\bar{g}_h(\abs{\mat{r}-\mat{r}'};n(\mat{r})\bigr) - 1\bigr\}}{\abs{\mat{r}-\mat{r}'}} \notag \\ &= \frac{1}{2} \iinteg{\mat{r}}{\mat{r}'}n(\mat{r}) \frac{n(\mat{r})\bigl\{\bar{g}_h\bigl(\abs{\mat{r}'};n(\mat{r})\bigr) - 1\bigr\}}{\abs{\mat{r}'}} \notag \\ &= \integ{\mat{r}}n(\mat{r})\varepsilon_{\text{xc}}\bigl(n(\mat{r})\bigr), \end{align} where the function $\varepsilon_{\text{xc}}(n)$ is just the exchange-correlation energy per electron of the homogeneous electron gas. This expression for the xc energy of the LDA shows explicitly that its hole is approximated by a spherical one. As mentioned above xc holes are not spherical~\cite{PhD-Buijse1991, BuijseBaerends2002}. The original weighted density approximation (WDA)~\cite{AlonsoGirifalco1978, GunnarssonJonsonLundqvist1979} improves on the shape of the xc hole by not replacing $n(\mat{r}')$ by $n(\mat{r})$ in the xc-energy~\eqref{eq:xcEnergy}. Since the sum-rule~\eqref{eq:xc holeSumRule} for the xc hole is no longer trivially satisfied, an effective density, $\bar{n}(\mat{r})$, is used as input into the pair-correlation function instead of the density at the reference position. The xc energy in the WDA is therefore \begin{align}\label{eq:WDAenergy} E_{\text{xc}}^{\text{WDA}} = \frac{1}{2} \iinteg{\mat{r}}{\mat{r}'}n(\mat{r}) \frac{n(\mat{r}')\bigl\{\bar{g}_h(\abs{\mat{r}-\mat{r}'};\bar{n}(\mat{r})\bigr) - 1\bigr\}}{\abs{\mat{r}-\mat{r}'}}, \end{align} where the effective density $\bar{n}(\mat{r})$ should be found by satisfying the sum-rule for the xc hole~\eqref{eq:xc holeSumRule} at every point~$\mat{r}$ \begin{align*} \integ{\mat{r}'} n(\mat{r}')\bigl\{\bar{g}_h(\abs{\mat{r}-\mat{r}'};\bar{n}(\mat{r})\bigr) - 1\bigr\} = -1. \end{align*} Unfortunately, it was found that the pair-correlation functional of the homogeneous electron gas is not a good approximation to the pair-correlation function of inhomogeneous systems as molecules and surfaces. Using a more localized function for $\bar{g} - 1$, the results were significantly improved~\cite{GunnarssonJones1980, SaddTeter1996, RushtonTozerClark2002}. From the expression for the WDA xc energy~\eqref{eq:WDAenergy} we immediately notice that the integral kernel is still asymmetric in the coordinates $\mat{r}$ and $\mat{r}'$ and therefore breaks the important symmetry $\bar{g}(\mat{r},\mat{r}') = \bar{g}(\mat{r}',\mat{r})$. This is not so important for the value of the xc energy. But for the corresponding xc potential, we obtain \begin{multline*} v_{\text{xc}}^{\text{WDA}}(\mat{r}) = \frac{1}{2} \integ{\mat{r}'} \frac{n(\mat{r}')\bigl\{\bar{g}_h(\abs{\mat{r}-\mat{r}'};\bar{n}(\mat{r})\bigr) - 1\bigr\}}{\abs{\mat{r}-\mat{r}'}} \\ {} + \frac{1}{2} \integ{\mat{r}'} \frac{n(\mat{r}')\bigl\{\bar{g}_h(\abs{\mat{r}-\mat{r}'};\bar{n}(\mat{r}')\bigr) - 1\bigr\}}{\abs{\mat{r}-\mat{r}'}} \\ {} + \frac{1}{2} \iinteg{\mat{r}'}{\mat{r}''}n(\mat{r}')\frac{n(\mat{r}'')}{\abs{\mat{r}'-\mat{r}''}} \\ {} \times \frac{\delta\bar{g}_h(\abs{\mat{r}'-\mat{r}''};\bar{n}\bigr)}{\delta \bar{n}(\mat{r}'')} \frac{\delta\bar{n}(\mat{r}'')}{\delta n(\mat{r})}. \end{multline*} The first term actually decays as $-1/(2r)$ as is obvious from the sum-rule. The long-range behavior of the other two terms is not so obvious, but in practice they turn out to decay exponentially~\cite{SaddTeter1996, RushtonTozerClark2002}. Therefore, the xc potential decays too slowly (as $-1/(2r)$) compared to the proper $-1/r$ decay~\cite{GunnarssonJonsonLundqvist1979}. The incorrect long-range behavior of the potential is expected to have a significant effect on properties which depend strongly on a proper xc potential such as the ionization energy and Rydberg excitations~\cite{AlmbladhBarth1985, LeeuwenGritsenkoBaerends1995, SchipperGritsenkoGisbergen2000}. This failure can be attributed to the broken symmetry of the pair-correlation function. Consequently, one of the requirements of our new functional will be that it should satisfy the proper symmetry of the pair-correlation function, i.e.\ $\bar{g}(\mat{r},\mat{r}') = \bar{g}(\mat{r}',\mat{r})$. \subsection{Step structure from exchange} As mentioned in the introduction, the steps in the KS potential are important for a proper description of the atomic shell structure~\cite{LeeuwenGritsenkoBaerends1995, LeeuwenGritsenkoBaerends1996} and also the neutral dissociation of hetero-nuclear molecules~\cite{PerdewParrLevyBalduz1982, GritsenkoBaerends1996}. It has been shown that the necessary steps for the atomic shell structure are already featured by the exchange energy~\cite{LeeuwenGritsenkoBaerends1995}. However, the necessary step in the dissociation of hetero diatomic molecules is a less clear case, since the spin-symmetry has to be broken to provide the necessary localization~\cite{FuksRubioMaitra2011, MakmalKummelKronik2011}. \begin{figure*}[t] \includegraphics[width=\textwidth]{holesEqui} \\ \caption{(Color online) The different holes of the H$_2$ molecule at $R_{\text{H--H}} = 1.4$ Bohr in atomic units and the reference electron at 0.3 Bohr to the left of the right nucleus along the bond axis ($\vecr_{\text{ref}} = (0,0,0.4)$ Bohr). The positions of the nuclei are indicated by the blue lines and the position of the reference electron is indicated by the red line. The left panel shows the exchange hole, $\bar{\rho}_{\text{x}}(\mat{r}|\vecr_{\text{ref}}) = -\abs{\sigma_g(\mat{r})}^2$, the middle panel shows the correlation hole, $\rho_{\text{c}}(\mat{r}|\vecr_{\text{ref}})$, which provides a small correction to have the more localized real hole, $\rho_{\text{xc}}(\mat{r}|\vecr_{\text{ref}})$.} \label{fig:H2holesEqui} \includegraphics[width=\textwidth]{holesFive} \caption{(Color online) Similar to the previous plots, but now for $R_{\text{H--H}} = 5.0$ Bohr. The reference electron is still at 0.3 Bohr to the left of the right nucleus along the bond axis ($\vecr_{\text{ref}} = (0,0,2.2)$ Bohr now). The exchange hole, $\bar{\rho}_{\text{x}}(\mat{r}|\vecr_{\text{ref}})$, remains completely delocalized, so requires an equally large correction from the correlation hole, $\rho_{\text{c}}(\mat{r}|\vecr_{\text{ref}})$, to obtain the real hole, $\rho_{\text{xc}}(\mat{r}|\vecr_{\text{ref}})$, localized around the reference electron.} \label{fig:H2holes5} \end{figure*} Closely related, the exchange energy also has the necessary features for the integer discontinuity~\cite{LeeuwenGritsenkoBaerends1995, GritsenkoBaerends1996}, since the exchange term often changes radically when crossing an integer number of electrons due to the usual idempotency of the KS density matrix. The corresponding hole, the exchange hole, can be expressed in spin-integrated quantities as \begin{align}\label{eq:xHoleDef} \bar{\rho}_{\text{x}}(\mat{r}|\vecr_{\text{ref}}) \equiv -\frac{1}{2}\frac{\abs{\gamma_s(\mat{r},\vecr_{\text{ref}})}^2}{n(\vecr_{\text{ref}})}, \end{align} where the spin-integrated KS density matrix is defined as \begin{align*} \gamma_s(\mat{r},\mat{r}') &\equiv \sum_{\sigma}\gamma(\mat{r}\sigma,\mat{r}'\sigma) \notag \\ &\equiv \sum_{\sigma}\brakket{\Psi_s}{\crea{\psi}(\mat{r}'\sigma')\anni{\psi}(\mat{r}\sigma)}{\Psi_s}, \end{align*} with $\Psi_s$ as the KS wavefunction. The KS density matrix can alternatively be expressed directly in terms of the KS orbitals $\phi_k(\mat{r})$ as \begin{align}\label{eq:KS1RDM} \gamma_s(\mat{r},\mat{r}') = \sum_kn_k\phi_k(\mat{r})\phi^*_k(\mat{r}'), \end{align} where $n_k$ are the occupation numbers, being simply 0 or 2 in the non-degenerate case. The exchange hole has the convenient property that it already satisfies the hole xc hole sum-rule~\eqref{eq:xc holeSumRule}. Since exchange already satisfies a number of important properties, it is often used as a starting point to model the full exchange-correlation effects. Traditionally, one adds a correction term, correlation, defined as the \emph{difference} between the exact exchange-correlation quantities and the ones with one exchange taken into account. For example the correlation hole is simply defined as \begin{align*} \bar{\rho}_{\text{c}}(\mat{r}|\vecr_{\text{ref}}) \equiv \bar{\rho}_{\text{xc}}(\mat{r}|\vecr_{\text{ref}}) - \bar{\rho}_{\text{x}}(\mat{r}|\vecr_{\text{ref}}). \end{align*} Although this approach has some appeal, it is inconvenient in practice, especially in bond-breaking situations. We show that explicitly in the next section. \subsection{Bond breaking} \begin{figure*}[t] \includegraphics[width=\textwidth]{ldaHolesFive} \caption{(Color online) LDA holes for the H$_2$ molecule at $R_{\text{H--H}} = 5.0$ Bohr in atomic units. The reference electron is again at 0.3 Bohr to the left of the right nucleus along the bond axis ($\vecr_{\text{ref}} = (0,0,2.2)$. Both the LDA x hole and c hole are localized, so do not resemble the exact ones. However, their sum, $\rho_{\text{xc}}(\mat{r}|\vecr_{\text{ref}})$, has much better resemblance to the full xc hole.} \label{fig:LDAholes5} \end{figure*} In Fig.~\ref{fig:H2holesEqui} we have plotted the different holes, $\bar{\rho}_{\text{x}}(\mat{r}|\vecr_{\text{ref}})$, $\rho_{\text{c}}(\mat{r}|\vecr_{\text{ref}})$ and $\rho_{\text{xc}}(\mat{r}|\vecr_{\text{ref}})$ for the H$_2$ molecule at equilibrium distance $R_{\text{H--H}} = R_{\text{e}} = 1.4$ Bohr. Note that the quantities with correlation are at full coupling strength ($\lambda = 1$) and not the integrated ones. Ideally we would like to have shown the integrated ones, but obtaining them is a rather difficult task. Since the effect of the kinetic energy is not very large on the total energy~\cite{PhD-Leeuwen1994}, we expect that the holes do not differ too much as well. The reference electron is fixed at 0.3 Bohr to the left from the right nucleus. The holes were calculated from full configurations interaction (CI) results using the 1s, 2s, 3s, 2p and 3d hydrogen wavefunctions on each atom as a basis set \footnote{The 3p orbitals caused linear dependency problems at $R_{\text{H--H}} = R_{\text{e}} = 1.4$ Bohr, so they could not be included.}. The exchange hole (x hole) can be worked out to be \begin{align*} \bar{\rho}_{\text{x}}(\mat{r}|\vecr_{\text{ref}}) = -2\frac{\abs{\sigma_g(\mat{r})}^2\abs{\sigma_g(\vecr_{\text{ref}})}^2}{n(\vecr_{\text{ref}})} = -\abs{\sigma_g(\mat{r})}^2, \end{align*} thus the exchange hole is actually independent of the position of the reference electron. However, the real hole is deeper around the reference electron and therefore, depending on the position of the reference electron, the correlation hole (c hole) has to add and remove an equal amount of the hole to deepen it around the reference electron. Although the xc hole is more localized around the reference position, it still shows the two-peak structure of the x hole. The localization effect becomes more prominent if we consider the hydrogen molecule at an elongated bond distance of $R_{\text{H--H}} = 5.0$ Bohr (Fig.~\ref{fig:H2holes5}). Again, the x hole is independent of the position of the reference electron and is completely delocalized over the molecule. However, the real hole is completely localized around the reference electron, which is again located at 0.3 Bohr to the left from the right nucleus. Therefore, the c hole has to completely remove the hole from the left side of the molecule at put it back at the right side such that it integrates still to $-1$ electron. The c hole can not be regarded as a ``small'' correction to the x hole anymore, since it is actually equal in magnitude. The lack of the ``small'' c hole in the Hartree--Fock (HF) description correcting the x hole is the main reason for the bad performance of HF for the dissociation of molecules. It is instructive to consider the LDA holes, since they explain why DFT and its predecessor, X$\alpha$~\cite{Slater1951}, are so successful. The LDA holes are actually the $\lambda$-integrated ones, so a direct comparison is strictly not correct. Luckily, however, in the dissociation limit the $\lambda$-integrated and the exact hole at $\lambda = 1$ are identical~\cite{LeeuwenGritsenkoBaerends1996, PerdewSavinBurke1995}. The reason is that at long bond distances the interaction term $\lambda/r_{12}$ for $\lambda > 0$ will favor wavefunction configurations in which the electrons are residing on different atoms, i.e.\ the Heitler--London wavefunction. The density corresponding to this ground state wavefunction $\Psi_{\lambda}$ will be exactly equal to the wavefunction $\Psi_1$ at full coupling strength. It thus follows that $\Psi_{\lambda} = \Psi$ for $\lambda >0$ at infinite bond distance. At $\lambda = 0$ the wavefunction simply remains a KS determinant with a doubly occupied $\sigma_g$ orbital. However, the $\lambda=0$ region becomes unimportant~\cite{TealeCorianiHelgaker2009} in the $\lambda$-integration for the pair-correlation function and hence $g = \bar{g}$. Therefore at the longer bond distance of $R_{\text{H--H}} = 5.0$ Bohr the exact xc hole should be close to its $\lambda$-integrated counterpart. In Fig.~\ref{fig:LDAholes5} we show the LDA holes for the hydrogen molecule at $R_{\text{H--H}} = 5.0$ Bohr. From our discussion in Sec.~\ref{sec:SymAndAsymp} it is clear what the definition of the xc hole of the LDA should be. The following expression~\cite{GunnarssonJonsonLundqvist1979, DreizlerGross1990} is consistent with the LDA energy expression~\eqref{eq:LDAenergy} \begin{align}\label{eq:LDAhole} \bar{\rho}^{\text{LDA}}_{\text{xc}}(\mat{r}|\mat{r}_{\text{ref}}) = n(\mat{r}_{\text{ref}}) \bigl(\bar{g}_h(n(\mat{r}_{\text{ref}}),\abs{\mat{r} - \mat{r}_{\text{ref}}}) - 1\bigr). \end{align} Gori--Giorgi and Perdew made an accurate model for the pair-correlation function of the homogeneous electron gas~\cite{Gori-GiorgiPerdew2002}, which we used to calculate $\bar{\rho}^{\text{LDA}}_{\text{xc}}$. The x hole can be calculated by using the exchange part of the electron gas pair-correlation function in this expression and the c hole is simply defined as the difference between the other two. The most striking feature of the LDA holes in Fig.~\ref{fig:LDAholes5} is that the x hole is localized instead of delocalized over the two atoms, just as is the case for the exact x hole. Although the LDA x hole does not resemble the exact x hole at all, the full xc hole (the one of interest) is actually modeled quite well. Especially, if we consider the exact x hole which is the hole employed in HF theory, the LDA xc hole provides a large improvement. We also see that the LDA c hole only provides a small correction to the x hole: it removes the outward oscillations and localizes the LDA hole a bit further. Since the correction from the c hole is so small, we can understand why the old X$\alpha$ method already outperformed HF so much. In particular, the deepening of the hole was empirically taken into account by scaling the exchange prefactor, $\alpha$, from its theoretical value, $2/3$, to $0.7$. These observations concerning the LDA holes also make it clear that it does not make sense at all to add ``exact exchange'' to LDA correlation. The same holds for GGA holes, since they are quite similar to LDA holes, only having slightly more wild oscillations and a discontinuity due to their cut-off to satisfy the sum-rule~\eqref{eq:xc holeSumRule} as additional features~\cite{SlametSahni1991, BurkePerdewWang1998}. \section{The screened exchange Model} \label{sec:SXmodel} Considering these facts about the x hole in dissociating H$_2$, it seems to be unwise to add a correlation hole to an exact exchange hole. It will be hard to build a model for the correlation hole with the proper strongly varying features. But we also know that exchange effects often dominate and that correlation effects only provide a modification of the former. This observation suggests that we should not add correlation to exchange but rather modify the shape of the x hole by some correlation factor. From the holes for the hydrogen molecule (Figs~\ref{fig:H2holesEqui} and~\ref{fig:H2holes5}) we observe that the main effect of correlation is to \emph{localize} the x hole. This is not special for the H$_2$ molecule, but is the main feature of correlation in any system, even in the electron gas where the correlation reduces the long range behavior of the x hole from $r^{-4}$ to $r^{-8}$~\cite{Gori-GiorgiSacchettiBachelet2000a, Gori-GiorgiSacchettiBachelet2000b}. A further example was provided long ago by J.C. Slater when he pointed out that atomic term energies were often accurately described by term dependent Hartree--Fock theory (``exact exchange '') by simply reducing Slater's $F_k$ and $G_k$ integrals by $\sim25\%$~\cite{Slater1960}. Following the discussion above it seems rather natural to use the following Ansatz for the xc energy $E_{\text{xc}}$ of an inhomogeneous system \begin{align}\label{eq:ExcSX} E_{\text{xc}} = -\frac{1}{4}\iinteg{\mat{r}}{\mat{r}'}\frac{\abs{\gamma_s(\mat{r},\mat{r}')}^2}{\abs{\mat{r}-\mat{r}'}}F(\mat{r},\mat{r}'). \end{align} Here, $\gamma_s(\mat{r},\mat{r}')$ is the spin-integrated non-interacting density matrix of the KS orbitals~\eqref{eq:KS1RDM} and the ``screening''-factor $F$ is intended to provide the necessary modification of exact exchange which will take care of the effects of correlations. The exact expression for $F$ is \begin{align*} F(\mat{r},\mat{r}') \equiv \frac{\bar{\rho}_{\text{xc}}(\mat{r}|\mat{r}')}{\bar{\rho}_{\text{x}}(\mat{r}|\mat{r}')} \end{align*} and by multiplying the numerator and denominator by $n(\mat{r})$ we immediately see that the screening function $F(\mat{r},\mat{r}')$ is symmetric in its arguments $\mat{r}$ and $\mat{r}'$, cf.~\eqref{eq:xcHoleDef} and~\eqref{eq:xHoleDef}. Our task is thus to find a reasonable model for $F$ and we stress again the importance of keeping the symmetry of $F(\mat{r},\mat{r}')$ in order to have an ensuing xc potential with the correct $-1/r$ tail outside finite systems. We also believe it to be essential to have a model which satisfies the sum-rule for the xc hole and in terms of $F$, our model should thus obey \begin{align*} \integ{\mat{r}'}\abs{\gamma_s(\mat{r},\mat{r}')}^2F(\mat{r},\mat{r}') = 2n(\mat{r}), \end{align*} If we, like the founding fathers (KS), first turn to the electron gas, we realize that $F$, and also $\gamma_s$ for that matter, must be spherical functions of only $\abs{\mat{r}-\mat{r}'}$. In the spirit of the older WDAs we could thus attempt such an Ansatz for our $F$ function. It is, however, important here to stress that in the original WDAs it is almost the entire xc hole which is modeled in this way, whereas in our case we just model a modification of the full, in general non-spherical x hole. As mentioned previously, we would also like to model the $F$-function in the case of the dissociation of H$_2$. In the dissociation limit of the hydrogen molecule the $F$-function actually takes the form \begin{align*} F_{\text{HL}}(\mat{r},\mat{r}') \approx f_a(\mat{r})f_a(\mat{r}') + f_b(\mat{r})f_b(\mat{r}') \end{align*} in terms of two one-point functions $f_a$ and $f_b$ located on the different hydrogen atoms. This follows from the fact that in the limit of large separation between the nuclei, the Heitler--London wavefunction becomes exact---but we will not present the details here. But it means that in this limit, the $F$-function is very far from spherical. When the two electrons are on different nuclei, the $F$-factor should vanish, because we are then dealing with two non-interacting subsystems and there is neither exchange nor correlation. When both electrons are on the same atom, the $F$-funcion should actually be 2 to make the xc hole equal to the negative of the local density. In this way the xc hole will precisely remove the full self-interaction on each atom---not just half the self-interaction as Hartree--Fock does---and we recover the correct limit of two separate and non-interacting atoms. As suggested by the above observations, the following Ansatz for the screening function $F(\mat{r},\mat{r}')$ might stand a chance of carrying us from the limit of a homogeneous system to that of the complete breaking up of the H$_2$ bond \begin{align}\label{eq:F_SX} F^{\text{SX}}(\mat{r},\mat{r}') \equiv A(\mat{r})A(\mat{r}')\, h\bigl(\abs{\mat{r}-\mat{r}'},\bar{n}(\mat{r},\mat{r}')\bigr). \end{align} Here, the spherical function $h$ has an effective screening radius $\bar{r}_s$ given by $4\pi\bar{r}_s^3 = 3/\bar{n}$. This form was also inspired by the success of a wavefunction for the helium atom by Hylleraas~\cite{Hylleraas1929} having precisely this form. We are then left with the choice of satisfying the hole sum-rule either by adjusting the one-point function $A(\mat{r})$, the hole-depth function, or by varying the effective density $\bar{n}(\mat{r},\mat{r}')$. We stress that the latter must be a symmetric two-point function in order not to jeopardize the asymptotics of the potential. In terms of the Ansatz~\eqref{eq:F_SX} the sum-rule reads \begin{align}\label{eq:Asumrule} A(\mat{r})\integ{\mat{r}'}\abs{\gamma_s(\mat{r},\mat{r}')}^2A(\mat{r}')\, h\bigl(\abs{\mat{r}-\mat{r}'},\bar{n}(\mat{r},\mat{r}')\bigr) = 2n(\mat{r}). \end{align} The effective density is expected to vary in a similar way as the density itself (from very small to very large values) and it is a two-point function. The hole-depth function $A$ on the other hand is a one-point function of limited variation---typically between zero and two depending on the normalization of the function $h$. Consequently, it appears to be numerically more stable to use the $A$-function for the purpose of satisfying the sum-rule. Indeed, we have encountered no difficulties in solving~\eqref{eq:Asumrule} in our applications. A further argument in favor of this choice is a lack of guidance from the electron gas when determining the $A$-factor, should we have chosen to satisfy the sum-rule by varying the effective density $\bar{n}$. Most WDA models have used the latter procedure, but again, they have not considered an $A$-factor. In our case we are thus free to choose the effective density. Typical choices are $\bar{n}(\mat{r},\mat{r}') = \frac{1}{2}\bigl(n(\mat{r}) + n(\mat{r}')\bigr)$ or $\bar{n}(\mat{r},\mat{r}') = n\bigl((\mat{r}+\mat{r}')/2\bigr)$. in previous attempts to construct a symmetric version of an WDA-like model we have found that the first of these suggestions gave rise to numerical difficulties, whereas Gunnarsson et al.~\cite{GunnarssonJonsonLundqvist1979} encountered difficulties with the second choice. A choice which seems to work in our previous attempts is \begin{align}\label{eq:avDensDef} \bar{n}(\mat{r},\mat{r}') \equiv \sqrt{n(\mat{r})n(\mat{r}')}, \end{align} which is the definition of the effective density $\bar{n}$ which we will use here. We stress, however, that there is no compelling reason for this choice. As a matter of fact, this is one part of our new model which we might have to revise in future attempts to improve the accuracy of the model. It is not clear what properties and shape the screening function $h$ should have. However, since its main task is to screen the x hole, we will use the very simple form inspired by the screened Coulomb (Yukawa) interaction \begin{align}\label{eq:HEGscreen} h_{\text{HEG}}(r, \bar{n}) \equiv \textrm{e}^{-D(\bar{n})r_{12}}. \end{align} The function $D(n)$ is fitted to the homogeneous electron gas such that it yields the exact xc energies for all homogeneous densities. In this way also the kinetic energy contribution to the xc energy is included. More details on the fit can be found in Appendix~\ref{ap:HEGscreen}. This form for the screening function is definitely too simplistic. More knowledge is required to build more accurate SX models. This will be the subject of future research. \section{Illustrative results} \label{sec:results} One of the most severe tests for our new SX functional will be if it performs well for dissociating molecules. To keep matters simple, we have limited ourselves to the hydrogen molecule. The most natural grid for calculations on a diatomic molecule is based on a prolate spheroidal coordinate system. A planar elliptical grid with foci at the two nuclei and $z$-axis joining those foci is rotated about that axis to generate the full grid. More details on the grid used are in Appendix~\ref{ap:grid}. As a further simplification we used the density from a full CI calculation in the same basis as used before (1s, 2s, 3s, 2p and 3d hydrogen wavefunctions). This CI expansion is not very good for obtaining an accurate total energy. However, we believe that the density will be accurate enough for the SX model. \begin{figure}[t] \includegraphics[width=\columnwidth]{SXenergy} \\ \caption{(Color online) Comparison of the total energy from the simple SX model with the ones from the full CI calculation for varying bond-length.} \label{fig:totalEnergies} \end{figure} \begin{figure*}[t] \includegraphics[width=\textwidth]{modelHolesEqui} \\ \caption{(Color online) The different model holes of the H$_2$ molecule at $R_{\text{H--H}} = 1.4$ Bohr in atomic units and the reference electron at 0.3 Bohr to the left of the right nucleus along the bond axis ($\vecr_{\text{ref}} = (0,0,0.4)$ Bohr). The positions of the nuclei are indicated by the blue lines and the position of the reference electron is indicated by the red line. The right most panel shows the exact (not integrated) xc hole for comparison. From left to right, the panels show the xc holes from the SX model with the $h_{\text{HEG}}$ screening function ~\eqref{eq:HEGscreen}, the SX model with the $h_1$ screening function~\eqref{eq:SlaterScreening}, the SX model with the $h_2$ screening function~\eqref{eq:GaussScreening} and the LDA hole evaluated as in~\eqref{eq:LDAhole} with the pair distribution from Ref.~\cite{Gori-GiorgiPerdew2002}.} \label{fig:modelHolesEqui} \includegraphics[width=\textwidth]{modelHolesFive} \caption{(Color online) Similar to the previous plots, but now for $R_{\text{H--H}} = 5.0$ Bohr. The reference electron is still at 0.3 Bohr to the left of the right nucleus along the bond axis ($\vecr_{\text{ref}} = (0,0,2.2)$ Bohr now).} \label{fig:modelHoles5} \end{figure*} The first step in evaluating the SX model is to solve for the hole-depth function, $A(\mat{r})$, by solving the integral equation~\eqref{eq:Asumrule} on the grid. Once the hole-depth function is obtained, we performed the double integral~\eqref{eq:ExcSX} with $F^{\text{SX}}$~\eqref{eq:F_SX} to obtain the xc energy according to our simple SX model. To calculate the total energy, we note that the non-interacting kinetic energy can be directly obtained from the density for the two-electron system, $T_s = \frac{1}{2}\int\bigl(\nabla\sqrt{n}\bigr)^2$, and the Hartree and nuclear contributions are already known from the full CI calculation. In Fig.~\ref{fig:totalEnergies} we compare the total energies from the SX model with the ones from a full CI calculation as a function of the distance, $R_{\text{H--H}}$, between the hydrogen atoms. As mentioned before the Slater basis is too poor to obtain a good total energy, so we performed a full CI calculation with the \textsc{dalton} package~\cite{Dalton} in an aug-cc-pVQZ basis~\cite{cc-pVT/QZ_H_B-Ne_aug_H} as a reference and additionally the corresponding HF result is shown as well. Unfortunately our SX model with the simplistic screening function is not performing much better than the HF method. It follows the HF curve rather closely and only around $R_{\text{H--H}} \approx 6$ Bohr does the curve start to bend downwards to the correct total energy. The most important feature of the SX model is that it directly models the xc hole and therefore we can look at it what is going wrong. In Fig.~\ref{fig:modelHolesEqui} in the left panel, we show the hole at $R_{\text{H--H}} = 1.4$ Bohr. If we compare with the exact holes in Fig.~\ref{fig:H2holesEqui}, we see immediately that our current SX model is not localizing the x hole sufficiently. Actually, the SX hole is almost identical to the x hole $\bar{\rho}_x$. In the left panel of Fig.~\ref{fig:modelHoles5} we show the hole the elongated distance $R_{\text{H--H}} = 5.0$ Bohr. Comparing with the exact holes in Fig.~\ref{fig:H2holes5} we see that the SX model actually does localize the x hole, but not sufficiently. The lack of localization of the hole explains exactly why the total energy in the SX model is consistently too high (Fig.~\ref{fig:totalEnergies}). Due to the $1/\abs{\mat{r}-\mat{r}'}$ term in the expression for the xc energy~\eqref{eq:xcEnergy}, a too delocalized hole does not stabilize the energy enough, which results in a too high energy. In retrospect we should not be surprised that parametrizing the screening function by the homogeneous electron gas did not work out. The main task of the screening function is to localize the hole which is most important in inhomogeneous systems like the hydrogen molecule. Therefore, its actual form and localization strength should be modeled by these systems and not the homogeneous electron gas. A detailed study and parametrization are beyond the aims of this article, but to reinforce our arguments for this statement, we also like to show some results for the following two heuristic screening functions \begin{subequations} \begin{align} \label{eq:SlaterScreening} h_1(r_{12},\bar{n}) &= \exp\biggl(-c_1\Bigl(\frac{r_{12}}{\bar{r}_s}\Bigr)\biggr), \\ \label{eq:GaussScreening} h_2(r_{12},\bar{n}) &= \exp\biggl(-c_2\Bigl(\frac{r_{12}}{\bar{r}_s}\Bigr)^2\biggr). \end{align} \end{subequations} The first one, $h_1$, keeps the Slater like form, but replaces the parametrization by the electron gas by a simple division by $\bar{r}_s$ to make the total dimensionless and a constant $c_1$ that we can choose to our liking. We found that $c_1 = 2.0$ gave a nice dissociation behavior for the energy. The second one is mainly included to emphasize that the required shape of the screening function is also unclear at the moment. Choosing the constant in the same manner as before, we found $c_2 = 0.5$ to be sufficient for our purposes. Note that both screening functions are not fitted to the electron gas anymore and are therefore not expected to give the correct xc energy density, $\varepsilon_{\text{xc}}(r_s)$, for the gas. The results for the energy from these screening functions are shown also in Fig.~\ref{fig:totalEnergies}. The situation is now rather different. The total energy is consistently underestimated. However, the shape of the curve is definitely an improvement. The total energy at elongated distances $R_{\text{H--H}} > 5$ Bohr remains rather constant as we choose the constants $c_i$ to do so. From the figure it is not immediately clear if the shape is also an improvement over the simple LDA functional. However, shifting the curves such that their minima coincide with the full CI minimum (Fig.~\ref{fig:shiftedEnergies}), we see that the energy from the SX models follow the full CI energy much closer. In particular the Gaussian, $h_2$ seems to reduce the overbinding most. Even at equilibrium distance the position and shape of the minimum seems to be somewhat improved. \begin{figure}[t] \includegraphics[width=\columnwidth]{ShiftedEnergy} \\ \caption{(Color online) Comparison of the total energy of the LDA functional~\cite{Gori-GiorgiPerdew2002} and the SX model with the Gaussian, $h_2$, as its screening function with the full CI results. The total energy of the LDA and SX model are shifted such that the minima coincide with the full CI value.} \label{fig:shiftedEnergies} \end{figure} Again, since we have built a direct model for the hole we can explain both features. Considering the holes from the heuristic $h_1$ and $h_2$ screening functions in Figs~\ref{fig:modelHolesEqui} and~\ref{fig:modelHoles5}, we see that both screening function are too powerful: they screen the x hole too much. This explains why the total energy is consistently lower than the full CI result: the hole becomes too compact. Since the hole is even more compact for $h_1$ screening function than the $h_2$ screening function, the energy of the $h_1$ screening function is even lower than the energy of the $h_2$ screening function. However, if we consider the shapes of the holes, then they are definitely an improvement over the previous screening function. Especially at the elongated bond distance $R_{\text{H--H}} = 5.0$ Bohr, we see that both heuristic screening functions nicely localize the hole completely at the correct side, which explains their improved energy in the dissociation limit over the erroneous $1/R_{\text{H--H}}$ of HF. The LDA hole is also shown in Figs~\ref{fig:modelHolesEqui} and~\ref{fig:modelHoles5}. Compared to the HF, cf.\ $\bar{\rho}_x$ in Fig.~\ref{fig:H2holesEqui} and~\ref{fig:H2holes5}, the LDA hole is a big improvement since it localizes around the reference electron. This improvement is also visible in the total energy where the LDA does not exhibit the erroneous $1/R_{\text{H--H}}$ behavior as HF does (Fig.~\ref{fig:totalEnergies}). However, the shape of the LDA hole is ridiculous. It is spherical by construction and it does not have the peaked features at the nuclei. Instead, the LDA hole attains its minimum at the position of the reference electron. The SX holes, especially with the $h_2$ screening function, are an improvement over both the HF hole and the LDA hole. It correctly localizes the hole around the reference electron and still retains the distinct peaked features of the hole. Hence, the binding curve is much improved over HF and LDA (see Fig.~\ref{fig:shiftedEnergies}). \section{Conclusion} \label{sec:conclusion} The aim of the present work is to construct a functional for the exchange-correlation (xc) energy of DFT which applies to such diverse systems as the electron gas and the dissociation of the hydrogen molecule (H$_2$). To this end we try to find a model for the xc hole of these and intermediate systems in real space. It has been known for long that the exchange (x) hole captures many important features of the full one. For instance, in atoms term energies are often well described by reducing the exchange effects by $25\%$ and in the electron gas correlations have the effect of reducing the range of the pure x hole from a $r^{-4}$ decay to a $r^{-8}$ decay~\cite{Gori-GiorgiSacchettiBachelet2000a, Gori-GiorgiSacchettiBachelet2000b}, $r$ being the distance from the the center of the hole. Thus unlike previous models that have sought to model the xc hole in real space, our present model aims at modifying the full x hole. Our investigations have shown that in the case of the dissociation of H$_2$ the xc hole is qualitatively different from the x hole. Therefore, it is an unwise strategy to add a correlation (c) hole to a full x hole. The former would have to replace the latter with a full xc hole with appropriate features. We show here that this can be achieved in a more natural manner by multiplying the x hole by an appropriate correlation factor. Judging from the electron gas, the correlation factor might be chosen as a function of the distance $\abs{\mat{r}-\vecr_{\text{ref}}}$ between an electron and a reference electron. (We here again remind the reader that the xc hole describes the depletion of negative charge around an electron known to be at the reference position $\vecr_{\text{ref}}$.) Unfortunately, our investigations have shown that such a simple correlation factor will have difficulties in moving half an x hole from one atom to the other---which is the appropriate effect of correlation in H$_2$ at large nuclear separation. Instead, we have chosen to include in our correlation factor an additional factor $A(\mat{r})$ depending on only one coordinate---a modulation of the depth of the xc hole. Such a factor is by symmetry just a constant in the electron gas and irrelevant to the shape of the hole in that case. Consequently, we have no guidance from the gas in choosing the $A$-factor. Instead, we have decided to determine this factor at every point in space from the sum-rule for the xc hole. This sum-rule is of course of utmost importance for obtaining an xc potential with the correct asymptotics outside finite systems ($-1/r$) from our model. This important property also requires the full symmetry in $\mat{r}$ and $\mat{r}'$ in the density multiplied xc hole, $n(\mat{r})\bar{\rho}_{\text{xc}}(\mat{r}'|\mat{r})$, something that we have emphasized throughout the paper. The decision to use the $A$-factor for satisfying the sum-rule leaves us with great freedom in choosing a screening factor depending only on $\abs{\mat{r}-\mat{r}'}$. Thinking about the electron gas it is natural to let this screening function have a screening length $\bar{r}_s$ determined by an effective density $\bar{n}$ according to $4\pi\bar{n}\bar{r}_s^3 = 3$. In the present work we have made the somewhat arbitrary choice $\bar{n}(\mat{r},\mat{r}') = \sqrt{n(\mat{r})n(\mat{r}')}$. We are, however, aware of that we might be forced to abandon this simple choice in future refinements of our model. For the actual shape of the screening function we have simply made a couple of reasonable choices for illustrational purposes. They are rapidly decaying, analytic functions with a few parameters with density dependence. One of the screening functions had its parameters specifically chosen such to reproduce the ``exact'' xc energies of the electron gas in the homogeneous limit, whereas two other screening functions had more \emph{ad-hoc} parameters to illustrate the effect of selecting different forms of the screening function. We could, however, also have attempted to find a screening function with a shape that would have allowed us to obtain accurate results for one- and two-electron systems. We would then have had an approximation which is able to properly dissociate H$_2$, which would be exact for the electron gas and quite accurate for the one- and two-electron cases. Such a functional is likely to give good total energies in a large number of systems. In the present work we have, however, refrained from going down this road. Instead, our aim here was to present the basic ideas and to leave further refinements to future investigations. In order to just illustrate our new approach, we thus chose to present results obtained from two simple, but rather arbitrary screening functions given by the equations~\eqref{eq:SlaterScreening} and~\eqref{eq:GaussScreening}. We have seen that the shorter ranged choice ($h_2$) gives better results, but they are still not very good. It is however, seen from both Fig.~\ref{fig:totalEnergies} and Fig.~\ref{fig:shiftedEnergies} that the errors are mainly associated with an inadequate description of a simple one-electron system. (At a bond distance of 10 Bohr we basically have two separate hydrogen atoms.) The most important result of the present work is that we managed to design a model which is able to describe the proper behavior of the xc hole of a hydrogen molecule as it dissociates. The details are not overly accurate, but we nurture real hope that they may fall in place by a better choice of the screening function. For now, however, we have left the search for such a function to future work.
\section{Introduction} By fundamental works of Nash \cite{na}, Gromov and Rokhlin \cite{gr} and G\"unther \cite{gu} we know that every $n$-dimensional smooth Riemannian manifold admits a smooth isometric embedding \footnote{We recall that an injective immersion is an embedding if it is a homeomorphism onto its image, by considering the image with the induced topology.} in an $N$-dimensional Euclidean space $\mathbb{R}^N$, for some $N\leq c(n)=\max\{n(n+5)/2, n(n+3)/2+5\}$. The estimate $c(n)$ was given by G\"unther \cite{gu}; Nash's and Gromov-Rokhlin's estimates are larger than this upper bound. Since then, the problem of finding the lowest possible codimension is one of the major open problems in the theory of isometric immersion. For books and surveys about this subject see Gromov and Rokhlin \cite{gr}, Jacobowitz \cite{jac}, Poznyak and Sokolov \cite{poso}, Aminov \cite{am}, Dajczer \cite{da}, Borisenko \cite{bo} and Han and Hong \cite{hh}. On the other hand, since the results of Nash, Gromov and Rokhlin and G\"unther follow as a consequence of existence theorems for certain PDE's, it is also an interesting problem to give the explicit construction of isometric immersions of a given Riemannian metric $M^n$ in $\mathbb{R}^m$, mainly if the attained codimension is strictly less than $c(n)-n$. This is the point of view of the present paper. Blanu\v sa \cite{bla, bla2} gave a method to construct injective smooth isometric immersions of the hyperbolic plane $\mathbb{H}^2$ in $\mathbb{R}^6$ and in the standard spherical space $\mathbb{S}^8$. Poznyak \cite{po} wrote about Blanu\v sa's surface: {\it{``There is no doubt that this result is one of the most elegant in the theory of immersion of two-dimensional manifolds in Euclidean space"}}. Blanu\v sa also constructed injective isometric immersions of $\mathbb{H}^n$ in $\mathbb{R}^{6n-5}$ and of an infinite M\"obius band with hyperbolic metric in $\mathbb{R}^8$ and in $\mathbb{S}^{10}$. His method was used and modified in further works: (i) Rozendorn \cite{rozen} constructed non-injective smooth isometric immersions of the plane $\mathbb{R}^2$ with the warped product metric of the form $d\sigma^2=dt^2+f(t)^2dx^2$ in $\mathbb{R}^5$ (this class of surfaces includes $\mathbb{H}^2$ with the metric $dt^2+e^{2t}dx^2$). Note that a celebrated theorem of Hilbert \cite{hilb} states that $\mathbb{H}^2$ cannot be isometrically immersed in $\mathbb{R}^3$. However, the existence of an isometric immersion of $\mathbb{H}^2$ in $\mathbb{R}^4$ or even an injective isometric immersion of $\mathbb{H}^2$ in $\mathbb{R}^5$ is still an open problem (a partial answer to the first problem was given by Sabitov \cite{sab}). (ii) Henke \cite{henke1, henke3} exhibited isometric immersions of $\mathbb{H}^n$ in $\mathbb{R}^{4n-3}$ and in $\mathbb{S}^{4n-3}$. (iii) Henke and Nettekoven \cite{henke2} showed that $\mathbb{H}^n$ can be isometrically embedded in $\mathbb{R}^{6n-6}$ whose image is the graph of a smooth map $g:\mathbb{R}^n\to \mathbb{R}^{5n-6}$. (iv) Azov \cite{azov} considered the space $\mathbb{R}^n=\mathbb{R}\times \mathbb{R}^{n-1}$ with one of the following metrics: $d\sigma^2 = dt^2+ f(t)^2\sum_{j=1}^{n-1}dx_j^2$ or $d\sigma^2 = g(x_1)^2\sum_{j=1}^n dx_j^2$ and constructed isometric immersions in $\mathbb{R}^{4n-3}$ and $\mathbb{S}^{4n-3}$. He also announced in \cite{azov2} the construction of isometric immersions of these classes of metrics in $\mathbb{R}^{4n-4}$ and $\mathbb{S}^{4n-4}$, if $n>2$. In this paper we deal with product manifolds $M^n=I\times \mathbb{R}^{n-1}$, where $I$ is an open interval, endowed with a multiple warped product metric of the form: \begin{eqnarray}\label{multiplewarped} d\sigma^2 = \,\rho(t)^2\,dt^2 +\eta_1(t)^2 dx_1^2+\ldots +\eta_{n-1}(t)^2 dx_{n-1}^2 \end{eqnarray} where $\rho(t),\eta_j(t)$, with $t\in I$ and $j=1,\ldots,n-1$, are positive smooth functions and $dx_1,\ldots, dx_{n-1}$ are the canonical coframes of $\mathbb{R}^{n-1}$. This class of metrics includes both Azov's metrics. We will modify Blanu\v sa's method to exhibit isometric immersions and, mainly, embeddings of this class of metrics in quadrics of semi-Euclidean spaces. It is worth to mention that, in general, the immersions obtained by Rozendorn, Henke and Azov are not injective. Based on this, we consider such embeddings the main contribution of the present work. There exists a wide literature about aspects of rigidity and nonimmersibility of these spaces (see for instance N\"olker \cite{no}, Chen \cite{ch}, Florit \cite{flo}, Dajczer and Tojeiro \cite{dato} and references therein). We recall that the semi-Euclidean space $\mathbb{R}^n_a$, with $a\in \{0,\ldots,n\}$, is simply the space $\mathbb{R}^n$ with the inner product of signature $(a,n-a)$ given by \begin{equation}\label{minkowiski} \left\langle\,,\right\rangle =\,-\,dx_1^2-\ldots - dx_a^2 + dx_{a+1}^2 + \ldots + dx_n^2, \end{equation} where $dx_j$, with $j=1,\ldots, n$, denote the canonical coframes of $\mathbb{R}^n$. For a given $c>0$, let $S^n_a(c)$ and $H^n_a(-c)$ be the following quadratic hypersurfaces (or, simply, quadrics): \begin{eqnarray*} &&\mathbb{S}^n_a(c) =\left\{x\in \mathbb{R}^{n+1}_a \bigm| \left\langle x,x\right\rangle={1}/{c}\right\};\\ &&\mathbb{H}^n_a(-c) = \left\{x\in \mathbb{R}^{n+1}_{a+1} \bigm| \left\langle x,x\right\rangle=-{1}/{c}\right\}. \end{eqnarray*} Both hypersurfaces are semi-Riemannian manifolds with signature $(a,n-a)$ and constant curvatures $c$ and $-c$, respectively. If $a=0$, then $\mathbb{S}^n_0(c)=\mathbb{S}^n(c)$ is the standard sphere and $\mathbb{H}^n_0(-c)=\mathbb{H}^n(-c)$ is the hyperbolic space. If $a=1$, the semi-Riemannian universal covering spaces of $\mathbb{S}^n_1(1)$ and $\mathbb{H}^n_1(-1)$ are called de Sitter $dS^n$ and anti-de Sitter $AdS^n$ spaces, respectively. Our main result is \begin{theorem}\label{mth} Let $M^n=I\times \mathbb{R}^{n-1}$ be as given in (\ref{multiplewarped}). Then, for all $c>0$ and $a\in \{0,\ldots, n-1\}$, the manifold $M^n$ admits: \begin{enumerate}[(i)] \item\label{euc-4n-2l-3} isometric immersions in $\mathbb{R}^{4n-3-2a}_{a}$, $\mathbb{S}^{4n-3-2a}_a(c)$, and $\mathbb{H}^{4n-3-a}_a(-c)$; \item\label{imb-8n-6a-7} isometric embeddings in $\mathbb{R}^{8n-7-6a}_a$, $\mathbb{S}^{8n-5-6a}_a(c)$ and $\mathbb{H}^{8n-7-5a}_a(-c)$. \end{enumerate} Moreover, all immersions and embeddings above are smooth and given explicitly. \end{theorem} In Remark \ref{non-inj} (see Section \ref{sec-proof-mth}), we observe that all immersions referred in Item {\it \ref{euc-4n-2l-3}} of Theorem \ref{mth} are not injective, provided that $a<n-1$. Based on Theorem \ref{mth}, it is natural to ask if every $n$-dimensional Riemannian manifold $M^n$ can be isometrically immersed in a semi-Euclidean $\mathbb{R}^N_a$ with $a>0$ and $N$ strictly less than the dimension $c(n)$ obtained by G\"unther \cite{gu}. As an application of Theorem \ref{mth} we generalize Rozendorn's surfaces \cite{rozen}. We have the following. \begin{corollary} Let $M^2=I\times \mathbb{R}$ be a warped product surface as given in (\ref{multiplewarped}). Then, for all $c>0$, the surface $M^2$ admits: \begin{enumerate} [(i)] \item non-injective isometric immersions in $\mathbb{R}^5$, $\mathbb{H}^5(-c)$, $\mathbb{S}^5(c)$ and $dS^3(c)$; \item isometric embeddings in $\mathbb{R}^9$, $\mathbb{H}^9(-c)$, $\mathbb{S}^{11}(c)$, $\mathbb{R}^3_1$, $AdS^4(-c)$ and $dS^5(c)$. \end{enumerate} Moreover, all immersions and embeddings above are smooth and given explicitly. \end{corollary} The space $Sol_3$ is a simply connected homogeneous $3$-dimensional space whose isometry group has dimension $3$. It is one of the eight models of the Thurston geometry and it can be viewed as $\mathbb{R}^3$ with the metric $ds^2=dt^2 + e^{2t}dx^2+e^{-2t}dy^2.$ It follows directly from Theorem \ref{mth} the following \begin{corollary}\label{two-dimensional} For all $c>0$, the space $Sol_3$ admits: \begin{enumerate}[(i)] \item non-injective isometric immersions in $\mathbb{R}^9$, $\mathbb{H}^9(-c)$, $\mathbb{S}^9(c)$, $\mathbb{R}^7_1$, $dS^7(c)$, $AdS^8(-c)$; \item isometric embeddings in $\mathbb{R}^{17}$, $\mathbb{H}^{17}(-c)$, $\mathbb{S}^{19}(c)$, $\mathbb{R}^{11}_1$, $AdS^{12}(-c)$ and $dS^{13}(c)$. \end{enumerate} Moreover, all immersions and embeddings above are smooth and given explicitly. \end{corollary} Let $f_l:M_l\to \mathbb{R}^{n_l}$, with $l=1,\ldots,k$, be smooth isometric immersions of the manifold $(M_l,g_l)$ in $\mathbb{R}^{n_l}$. Let $I$ be an open interval and $\rho(t),\eta_l(t)$, with $t\in I$ and $l=1,\ldots,k$, positive smooth functions. It is simple to show that the product manifold $M=I\times M_1\times\ldots\times M_k$ with the warped product metric \begin{equation}\label{warper metr riem} g=\rho(t)^2 dt^2 + \eta_1(t)^2 g_1+\ldots+\eta_k(t)^2 g_k \end{equation} can be isometrically immersed in $I\times \mathbb{R}^{n_1}\times\ldots\times\mathbb{R}^{n_k}$ with the metric \begin{equation*} d\sigma^2 = \rho(t)^2 dt^2 + \eta_1(t)^2 \delta_1+\ldots+\eta_k(t)^2\delta_k, \end{equation*} where $\delta_l$ denotes the Euclidean metric of $\mathbb{R}^{n_l}$. Thus it follows as a consequence of Theorem \ref{mth} the following result. \begin{corollary} With the notations being as above, we consider $n=n_1+\ldots+n_k$ and $a\in \{0,\ldots,n\}$. For all $c>0$, the manifold $M$ admits: \begin{enumerate}[(i)] \item isometric immersions in $\mathbb{R}^{4n+1-2a}_{a}$, $\mathbb{S}^{4n+1-2a}_{a}(c)$ and $\mathbb{H}^{4n+1-a}_{a}(-c)$; \item isometric embeddings in $\mathbb{R}^{8n+1-6a}_{a}$, $\mathbb{S}^{8n+3-6a}_{a}(c)$ and $\mathbb{H}^{8n+1-5a}_{a}(-c)$, provided that each $f_l$ is an embedding. \end{enumerate} \end{corollary} We would like to thank the referee for carefully reading the first version of this manuscript, pointing out mistakes which helped us to improve the manuscript. \section{Preliminaries.}\label{method} We recall Blanu\v sa's functions $\hat\psi_1,\hat\psi_2:\mathbb{R}\to \mathbb{R}$ defined by \begin{equation*} \hat\psi_1(u)=\sqrt{\frac{1}{A}\int_0^{u+1}\xi(\tau)d\tau} \ \ \mbox{ and } \ \ \hat\psi_2(u)=\sqrt{\frac{1}{A}\int_0^{u}\xi(\tau)d\tau}, \end{equation*} where $A=\int_0^{1}\xi(\tau)d\tau$ and $\xi(u)=\sin(\pi u)e^{\frac{-1}{(\sin(\pi u))^2}}$, if $u\in \mathbb{R}\setminus \Bbb Z$, and $\xi(u)=0$, if $u\in \Bbb Z$. Blanu\v sa proved in \cite{bla} that these functions are smooth, non-negative and satisfy: \begin{enumerate}[(a)] \item\label{bla-it1} $\hat\psi_j$ is periodic with period 2, for all $j=1,2$; \item\label{bla-it2} $\hat\psi_1^2+\hat\psi_2^2=1$, everywhere; \item\label{bla-it3} all the derivatives $\hat\psi_1^{(k)}(2l+1)=\hat\psi_2^{(k)}(2l)=0$, for all $l\in \Bbb Z$. \end{enumerate} The next two lemmas will be useful to prove Theorem \ref{mth}. They are simple consequences of Items \ref{bla-it1}, \ref{bla-it2} and \ref{bla-it3} above. To state them, let $I$ be an open interval and $\gamma:I\to \mathbb{R}$ an increasing smooth diffeomorphism. Consider the sequence $\mathbf{t}_k=\gamma^{-1}(k)$, with $k\in \Bbb Z$. The first lemma says the following. \begin{lemma}\label{bla-fun-I} The functions $\psi_j=\hat\psi_j\circ \gamma:I\to \mathbb{R}$ are smooth, non-negative and satisfy the following properties: \begin{equation}\label{bla-fun} \left\{ \begin{array}{l} \psi_1(t)^2+\psi_2(t)^2=1, \mbox{ everywhere in } I;\\ \psi_j\big((\gamma)^{-1}(u)\big)=\psi_j\big((\gamma)^{-1}(u+2)\big), \mbox{ for all } u\in \mathbb{R} \mbox{ and } j=1,2;\\ \psi_1^{(k)}(\mathbf{t}_{2l+1})=\psi_2^{(k)}(\mathbf{t}_{2l})=0, \mbox{ for all }k\geq 0 \mbox{ and } \mbox{ integers } l. \end{array}\right. \end{equation} \end{lemma} Let $S_1,S_2:I\to (0,\infty)$ be any positive step functions satisfying \begin{equation}\label{stp-fun}\left\{ \begin{array}{l} S_1 \mbox{ is constant on each interval } [\mathbf{t}_{2l+1},\mathbf{t}_{2l+3});\\ S_2 \mbox{ is constant on each interval } [\mathbf{t}_{2l}, \mathbf{t}_{2l+2}); \end{array}\right. \end{equation} for each integer $l$. The second lemma follows easily from Lemma \ref{bla-fun-I}. \begin{lemma}\label{lemmaderiv} For any $\eta\in C^\infty(I)$, the functions $\frac{\eta(t)\psi_j(t)}{S_j(t)}$ with $t\in I$ and $j=1,2,$ are smooth and their derivatives satisfy \begin{equation*} \frac{d^{\,k}}{dt^k}\left(\frac{\eta(t)\psi_j(t)}{S_j(t)}\right)=\frac{\frac{d^{\,k}}{dt^k}\Big(\eta(t)\psi_j(t)\Big)}{S_j(t)}, \end{equation*} for all integers $k\geq 0$. \end{lemma} \section{Proof of Theorem \ref{mth}}\label{sec-proof-mth} First we consider the map $\eta(t)=(\eta_1(t),\ldots,\eta_{n-1}(t))$, with $t\in I$, where each function $\eta_j$ is being as in (\ref{multiplewarped}). Consider the map $h:\mathbb{R}\to \mathbb{R}^2_1$ given by $h(u)=(\cosh(u),\sinh(u))$. Consider also the map $\varphi=(\varphi_1,\varphi_2):I\times \mathbb{R} \to \mathbb{R}^4$ where each map $\varphi_j:I\times \mathbb{R}\to \mathbb{R}^2$, with $j=1,2$, is given by \begin{equation}\label{map-euc} \varphi_j(t,u)=\frac{\psi_j(t)}{S_j(t)}\Big(\cos(S_j(t)u),\sin(S_j(t)u)\Big). \end{equation} The map $\varphi$ is introduced in \cite{henke2} for the case that $I=\mathbb{R}$ and $\gamma$ is the identity function. By using Lemma \ref{lemmaderiv}, we obtain \begin{eqnarray}\label{prelim} &&\frac{\partial(\eta_k(t) h(u))}{\partial t} = \eta_k'(t)\big(\cosh(u),\sinh(u)\big);\\ &&\frac{\partial(\eta_k(t) h(u))}{\partial u} = \eta_k(t)\big(\sinh(u),\cosh(u)\big); \nonumber\\ &&\frac{\partial (\eta_{k}(t)\varphi_j(t,u))}{\partial t} = \frac{(\eta_{k}(t)\psi_j(t))'}{S_j(t)}\Big(\cos(S_j(t)u),\sin(S_j(t)u)\Big); \nonumber\\ &&\frac{\partial (\eta_{k}(t)\varphi_j(t,u))}{\partial u} = \eta_{k}(t)\psi_j(t)\Big(-\sin(S_j(t)u)\,,\,\cos(S_j(t)u)\Big);\nonumber \end{eqnarray} for all $j=1,2$, $t\in I$, $u\in \mathbb{R}$ and $k=1,\ldots,n-1$. Now, set $a\in \{0,\ldots, n-1\}$ and let $b=n-1-a$. First we consider $b>0$. We will see that the case $b=0$ is easier. We write the semi-Euclidean space $\mathbb{R}^{4n-4-2a}_a=\mathbb{R}^{2a+4b}_{a}$ isometrically as the following form \begin{equation*} \mathbb{R}^{2a+4b}_{a}=(\mathbb{R}^{2}_1)^a\times \mathbb{R}^{4b} = \underbrace{\mathbb{R}^2_1 \times \ldots \times \mathbb{R}^2_1}_{\textrm{$a$ times }} \, \times \, \mathbb{R}^{4b}. \end{equation*} We denote by $x=(x_1,\ldots,x_{a+b})$ the coordinates of $\mathbb{R}^{n-1}=\mathbb{R}^{a+b}$. Let $P_1:\mathbb{R}^{a+b}\to \mathbb{R}^a$ and $P_2:\mathbb{R}^{a+b}\to \mathbb{R}^b$ be the standard orthogonal projections \begin{eqnarray*} &&\tilde x=P_1(x_1,\ldots,x_{a+b}) = (x_1,\ldots,x_a)\\ &&\bar x= P_2(x_1,\ldots,x_{a+b})=(x_{a+1},\ldots,x_{a+b}). \end{eqnarray*} Consider the maps \begin{equation}\label{ti-bar} \tilde\eta(t)=P_1(\eta(t)) \,\mbox{ and }\,\bar \eta(t)=P_2(\eta(t)), \end{equation} with $t\in I$ and let $\tilde\eta\star h:I\times \mathbb{R}^{n-1}\to (\mathbb{R}^2_1)^a$ and $\bar\eta\star\varphi:I\times \mathbb{R}^{n-1}\to \mathbb{R}^{4b}$ be the maps given by \begin{eqnarray} &&(\tilde\eta\star h)(t,x_1,\ldots,x_{a+b}) = \big(\eta_1(t)h(x_1),\ldots, \eta_a(t) h(x_a)\big)\in (\mathbb{R}^2_1)^a;\label{starfunction}\\ &&(\bar\eta\star\varphi)(t,x_{1},\ldots,x_{a+b}) = \big(\eta_{a+1}(t)\varphi(t,x_{a+1}), \ldots, \eta_{a+b}(t)\varphi(t,x_{a+b})\big)\in\mathbb{R}^{4b}\nonumber. \end{eqnarray} Since $h(u)\in \mathbb{R}^2_1$ and $\varphi(t,u)\in \mathbb{R}^4$, by using (\ref{prelim}), the pull-back symmetric tensors by $\tilde\eta\star h$ and $\bar\eta\star\varphi$ must satisfy \begin{eqnarray}\label{etah} (\tilde\eta\star h)^*(\left\langle\,,\right\rangle)&=& -|\tilde\eta'(t)|^2\,dt^2 + \eta_1(t)^2 dx_1^2+\ldots+ \eta_a(t)^2 dx_a^2;\\ (\bar\eta\star\varphi)^*(\left\langle\,,\right\rangle) &=& \epsilon(t)^2\,dt^2 + \eta_{a+1}(t)^2 dx_{a+1}^2 + \ldots + \eta_{a+b}(t)^2 dx_{a+b}^2, \nonumber \end{eqnarray} where $|\tilde\eta'(t)|^2=\eta_1'(t)^2+\ldots+\eta_a'(t)^2$ and $\epsilon:I\to [0,\infty)$ is given by \begin{equation}\label{ep_r} \epsilon(t)^2= \sum_{r=a+1}^{a+b}\left[\frac{\left(\left(\eta_{r}(t)\psi_1(t)\right)'\right)^2}{S_1(t)^2} + \frac{\left(\left(\eta_{r}(t)\psi_2(t)\right)'\right)^2}{S_2(t)^2}\right]. \end{equation} For the step functions $S_1,S_2:I\to (0,\infty)$ defined as in (\ref{stp-fun}), we can choose the steps $S_1|_{[\mathbf{t}_{2l+1}, \mathbf{t}_{2l+3})}$ and $S_2|_{[\mathbf{t}_{2l},\mathbf{t}_{2l+2})}$, with integer $l$, sufficiently large so that, for all $r=a+1,\ldots,a+b$, it holds \begin{eqnarray}\label{cond-ep} \big(\left(\eta_{r}(t)\psi_j(t)\right)'\big)^2 <\frac{1}{4b}S_j(t)^2 \,\rho(t)^2, \end{eqnarray} for all $t\in I$ and $j=1,2$. We obtain $\rho(t)^2-\epsilon(t)^2\geq \rho(t)^2-2\epsilon(t)^2> 0,$ for all $t\in I$. Let $f:I\times \mathbb{R}^{n-1}\to \mathbb{R}^{4n-3-2a}_a=\mathbb{R}\times(\mathbb{R}^2_1)^a\times \mathbb{R}^{4b}$ be the map \begin{equation*} f(t,x)= \left(\int_{t_0}^t \sqrt{\rho(\tau)^2+|\tilde\eta'(\tau)|^2-\epsilon(\tau)^2}d\tau\,,\, \tilde\eta\star h\,(t,x)\,,\,\bar\eta\star \varphi\,(t,x)\right). \end{equation*} If $b=0$, we define $f(t,x)$ simply by omitting $\epsilon(t)$ and $\bar\eta\star \varphi(t,x)$ in the expression of $f(t,x)$ above. By using (\ref{etah}), we obtain \begin{eqnarray*} f^*(\left\langle\,,\right\rangle) &=& (\rho(t)^2+|\tilde\eta'(t)|^2-\epsilon(t)^2) dt^2 + (\tilde\eta\star h)^*(\left\langle\,,\right\rangle) + (\bar\eta\star \varphi)^*(\left\langle\,,\right\rangle) \\ &=&\rho(t)^2 dt^2 + \eta_1(t)^2dx_1^2+\ldots + \eta_{a+b}(t)^2 dx_{a+b}^2\\ &=& d\sigma^2. \end{eqnarray*} This implies that $f:M^n\to \mathbb{R}^{4n-3-2a}_a$ is a smooth isometric immersion. We fix $c>0$. First we assume $b>0$. We choose the step functions $S_1,S_2$ sufficiently large so that (\ref{cond-ep}) holds. Let $\alpha:I\to [0,\infty)$ be the function given by \begin{equation}\label{def-a} \alpha(t) = \sum_{r=a+1}^{a+b}\left(\frac{\eta_r(t)^2\psi_1(t)^2}{S_1(t)^2}+\frac{\eta_r(t)^2\psi_2(t)^2}{S_2(t)^2}\right). \end{equation} Note that $\alpha(t)=\left\langle \bar\eta\star\varphi\,(t,x), \bar\eta\star\varphi\, (t,x)\right\rangle$. Let $f_h:I\times \mathbb{R}^{n-1}\to \mathbb{R}^{4n-2-a}_{a+1}=\mathbb{R}^2_1\times \mathbb{R}^a\times(\mathbb{R}^2_1)^a\times\mathbb{R}^{4b}$ be the map \begin{equation*} f_h(t,x) = \left(\sqrt{{1}/{c}+\alpha(t)}\,h(\theta_h(t))\,, \tilde\eta(t)\,, \tilde\eta\star h(t,x)\,, \bar \eta\star \varphi(t,x) \right), \end{equation*} where $\theta_h:I\to \mathbb{R}$ is the function defined by \begin{equation*} \theta_h(t) = \int_{t_0}^t \sqrt{\frac{1}{\frac{1}{c}+\alpha(\tau)}\left[\rho(\tau)^2-\epsilon(\tau)^2 + \frac{\alpha'(\tau)^2}{4\left(\frac{1}{c}+\alpha(\tau)\right)}\right]}d\tau. \end{equation*} If $b=0$, we define $f_h(t,x)$ simply by omitting $\epsilon(t)$, $\alpha(t)$ and $\bar\eta\star\varphi(t,x)$ in the expressions of $\theta_h(t)$ and $f_h(t,x)$ above. By (\ref{cond-ep}), we have $\rho(t)^2-\epsilon(t)^2>0$. Thus, in both cases $b=0$ and $b>0$, we have that $\theta_h$ is well defined, smooth and increasing. It is easy to see that $\left\langle f_h(t,x),f_h(t,x)\right\rangle = -\big({1}/{c}+\alpha(t) \big) + |\tilde\eta(t)|^2 -|\tilde\eta(t)|^2 + \alpha(t)= - {1}/{c},$ hence the image of $f_h$ is contained in $\mathbb{H}^{4n-3-a}_a(-c)$. By using (\ref{etah}), \begin{eqnarray*} f_h^*(\left\langle\,,\right\rangle) &=&\left[-\frac{1}{4}\left(\frac{\alpha'(t)^2}{\frac{1}{c}+\alpha(t)}\right) + \left(\frac{1}{c}+\alpha(t)\right)\theta'_h(t)^2 + |\tilde\eta'(t)|^2\right] dt^2 \\&& + (\tilde\eta\star h)^*(\left\langle\,,\right\rangle) + (\bar\eta\star \varphi)^*(\left\langle\,,\right\rangle)\\ &=& \rho(t)^2 dt^2 + \eta_1(t)^2 dx_1^2 + \ldots + \eta_{a+b}(t)^2 dx_{a+b}^2. \end{eqnarray*} This implies that $f_h:M^n\to \mathbb{H}^{4n-3-a}_a(-c)$ is an isometric immersion. Now, choose the step functions $S_1,S_2$ sufficiently large so that (\ref{cond-ep}) is satisfied and, moreover, for all $r=a+1,\ldots,a+b$, it holds \begin{eqnarray}\label{gransphere} \eta_r(t)^2\psi_j(t)^2<\frac{1}{8bc}S_j(t)^2, \end{eqnarray} for all $t\in I$ and $j=1,2$. By (\ref{def-a}) and (\ref{gransphere}), we obtain $0\leq \alpha(t) < \frac{1}{4c}$, for all $t\in I$. Let $f_s:I\times \mathbb{R}^{n-1}\to \mathbb{R}_a^{4n-2-2a}=\mathbb{R}^2\times (\mathbb{R}^2_1)^a\times \mathbb{R}^{4b}$ be the map defined by \begin{equation*} f_{s}(t,x) = \left(\sqrt{{1}/{c}-\beta(t)}\,g(\theta_s(t))\,, \, \tilde\eta\star h\,(t,x)\,,\,\bar\eta\star \varphi\,(t,x)\right), \end{equation*} where $\beta(t)=\alpha(t)-|\tilde\eta(t)|^2$, with $t\in I$, $g(u)=(\cos(u),\sin(u))$, with $u\in \mathbb{R}$, and $\theta_{s}:I\to\mathbb{R}$ is the function given by \begin{equation*} \theta_s(t)= \int_{t_0}^t \sqrt{\frac{1}{\frac{1}{c}-\beta(\tau)}\left[\rho(\tau)^2+|\tilde\eta'(\tau)|^2-\epsilon(\tau)^2- \frac{\beta'(\tau)^2}{4\big(\frac{1}{c}-\beta(\tau)\big)}\right]}\,d\tau. \end{equation*} If $b=0$, we define $f_s(t,x)$ by omitting $\epsilon(t)$, $\alpha(t)$ and $\bar\eta\star\varphi(t,x)$ in the expressions of $\theta_s(t)$ and $f_s(t,x)$ above. \begin{claim}\label{welldef} We can choose steps functions $S_1,S_2$, sufficiently large so that the function $\theta_s$ is well defined and smooth. \end{claim} In fact, first we assume $b=0$. In this case, by definition, we have $\beta(t)=-|\tilde\eta(t)|^2$. Hence, $\frac{1}{c}-\beta(t)>|\tilde\eta(t)|^2$. Moreover, $\beta'(t)^2 = 4\left\langle \tilde\eta'(t),\tilde\eta(t)\right\rangle^2\leq 4|\tilde\eta'(t)|^2|\tilde\eta(t)|^2$. Thus, \begin{equation*} \frac{\beta'(t)^2}{4\big(\frac{1}{c}-\beta(t)\big)}\le |\tilde\eta'(t)|^2. \end{equation*} By (\ref{cond-ep}), we have $\rho(t)^2-\epsilon(t)^2>0$. Thus we conclude that $\theta_s$ is well defined and smooth. Now, assume $b>0$. By (\ref{gransphere}), it holds $\frac{1}{c}-\beta(t)\ge \frac{1}{c}-\alpha(t)>0$. Using Lemma \ref{lemmaderiv}, we have \begin{equation}\label{deriv-a} \alpha'(t) = \sum_{r=a+1}^{a+b}\left[\frac{\left(\eta_r(t)^2\psi_1(t)^2\right)'}{S_1(t)^2}+\frac{\left(\eta_r(t)^2\psi_2(t)^2\right)'}{S_2(t)^2}\right]. \end{equation} Since $\beta'(t)=\alpha'(t)-2\left\langle \tilde\eta'(t),\tilde\eta(t)\right\rangle$, we obtain $\frac{\beta'(t)^2}{4} \leq \delta(t) + |\tilde \eta'(t)|^2|\tilde\eta(t)|^2,$ where $\delta:I\to [0,\infty)$ is the continuous function given by \begin{equation}\label{def-delta} \delta(t) = \left|\frac{\alpha'(t)^2}{4}- \alpha'(t)\left\langle\tilde\eta'(t),\tilde\eta(t)\right\rangle \right|. \end{equation} Using that $\frac{1}{c}-\beta(t)=\frac{1}{c}-\alpha(t)+|\tilde\eta(t)|^2>\frac{1}{2c}+|\tilde\eta(t)|^2$, it holds that \begin{equation*} \frac{\beta'(t)^2}{4(\frac{1}{c}-\beta(t))} \leq \frac{1}{\frac{1}{2c}+|\tilde\eta(t)|^2}(\delta(t) + |\tilde\eta(t)|^2|\tilde\eta'(t)|^2). \end{equation*} This implies that \begin{eqnarray*} \rho(t)^2+|\tilde\eta'(t)|^2-\epsilon(t)^2-\frac{\beta'(t)^2}{4(\frac{1}{c}-\beta(t))} &\geq& \rho(t)^2 - \epsilon(t)^2+|\tilde\eta'(t)|^2\left(1-\frac{|\tilde\eta(t)|^2}{\frac{1}{2c}+|\tilde\eta(t)|^2}\right) \\&& - \, \frac{\delta(t)}{\frac{1}{2c}+|\tilde\eta(t)|^2}\\ &=& \frac{1}{\frac{1}{2c}+|\tilde\eta(t)|^2}(\Gamma(t)-\delta(t)), \end{eqnarray*} where $\Gamma:I\to \mathbb{R}$ is the continuous function given by \begin{equation*} \begin{array}{l} \Gamma(t) = (\frac{1}{2c}+|\tilde\eta(t)|^2)\left(\rho(t)^2 - \epsilon(t)^2+|\tilde\eta'(t)|^2\left(1-\frac{|\tilde\eta(t)|^2}{\frac{1}{2c}+|\tilde\eta(t)|^2}\right)\right). \end{array} \end{equation*} Since the step functions $S_1$ and $S_2$ satisfy (\ref{cond-ep}), we obtain $\Gamma(t)>0$, for all $t\in I$. Furthermore, if $S_1(t)$ and $S_2(t)$ become larger, then $\Gamma(t)>0$ increases and $\delta(t)$ is as smaller as we want. So, we choose each step of $S_1$ and $S_2$ sufficiently large so that $\delta(t)<\Gamma(t)$, for all $t\in I$. This implies that \begin{equation}\label{well-def-sphere} \rho(t)^2+|\tilde\eta'(t)|^2-\epsilon(t)^2-\frac{\beta'(t)^2}{4(\frac{1}{c}-\beta(t))}>0, \end{equation} for all $t\in I$. Claim \ref{welldef} is proved. \\ It is easy to see that $\left\langle f_s(t,x),f_s(t,x)\right\rangle = {1}/{c}$, hence the image of $f_s$ is contained in $\mathbb{S}^{4n-3-2a}_a(c)$. By using (\ref{etah}), \begin{eqnarray*} f_s^*(\left\langle\,,\right\rangle) &=& \left(\frac{\beta'(t)^2}{4\left(\frac{1}{c}-\beta(t)\right)}+\left(\frac{1}{c}-\beta(t)\right)\theta'_s(t)^2\right)dt^2 + (\tilde\eta\star h)^*(\left\langle\,,\right\rangle) + (\bar\eta\star \varphi)^*(\left\langle\,,\right\rangle)\\ &=& \rho(t)^2 dt^2 + \eta_1(t)^2 dx_1^2 + \ldots + \eta_{a+b}(t)^2 dx_{a+b}^2. \end{eqnarray*} This implies that $f_s:M^n \to \mathbb{S}_a^{4n-3-2a}(c)$ is an isometric immersion. Item \ref{euc-4n-2l-3} is proved. \begin{remark}\label{non-inj} {\it The immersions $f$, $f_h$ and $f_s$ are not injective, if $b>0$. In fact, we take $t=\mathbf{t}_{2k}$, for some integer $k$. Let $x^1=(x^1_1,\ldots,x^1_{a+b})$ and $x^2 =(x^2_1,\ldots,x^2_{a+b})$ be vectors satisfying the following. \begin{enumerate}[(i)] \item $(x^1_1,\ldots,x^1_a) = (x^2_1,\ldots,x^2_a)$; \item $S_1(t)x^1_r=S_1(t)x^2_r + 2\pi l_{r}$, for some integer $l_r$, with $r=a+1,\ldots,a+b$ and $l_r\neq 0$ for some $r$. \end{enumerate} Notice that $\tilde\eta\star h\,(t,x^1) = \tilde\eta\star h\,(t,x^2)$, since $(x^1_1,\ldots,x^1_a) = (x^2_1,\ldots,x^2_a)$. Since $\psi_2(t)=\psi_2(\mathbf{t}_{2k})=0$ and $(\cos(S_1(t)x^1_r),\sin(S_1(t)x^1_r)) = (\cos(S_1(t)x^2_r),\sin(S_1(t)x^2_r))$, we obtain \begin{equation*} \psi_j(t)(\cos(S_j(t)x^1_r),\sin(S_j(t)x^1_r)) = \psi_j(t)(\cos(S_j(t)x^2_r),\sin(S_j(t)x^2_r)). \end{equation*} This implies that $\bar\eta\star\varphi(t,x^1)=\bar\eta\star\varphi(t,x^2)$. Since the first coordinates of $f, f_h$ and $f_s$ depend only on $t$, it follows that $f(t,x^1)=f(t,x^2)$, $f_h(t,x^1)=f_h(t,x^2)$, and $f_s(t,x^1)=f_s(t,x^2)$. Thus, the immersions $f, f_h$ and $f_s$ are not injective.} \end{remark} Now we will prove Item \ref{imb-8n-6a-7}. We will continue to assume the notations being as given in the proof of Item \ref{euc-4n-2l-3}. Let $T_1:\mathbb{R}\to (0,\frac{\pi}{2})$ and $T_2:\mathbb{R}\to \mathbb{R}$ be the smooth functions \begin{equation} \label{T1T2} T_1(u) = \frac{\pi}{4}\left(1+\tanh(u)\right) \ \ \mbox{ and } \ \ T_2(u) = \int_0^u \sqrt{1-T'_1(\tau)^2}d\tau. \end{equation} Note that $T_2$ is smooth since $T_1$ is analytic and $T_1'(u)=\frac{\pi}{4}\mathrm{sech}^2(u) \leq \frac{\pi}{4}<1$. Consider the map $\hat\varphi=(\varphi_{11},\varphi_{21},\varphi_{12},\varphi_{22}):I\times \mathbb{R}\to \mathbb{R}^8$, where each map $\varphi_{ji}:I\times \mathbb{R}\to \mathbb{R}^2$, with $i,j=1,2$, is defined by \begin{equation}\label{def-vp-jk} \varphi_{ji}(t,u) = \frac{\psi_j(t)}{S_j(t)}\Big(\cos\left(T_i(S_j(t)u)\right),\,\sin\left(T_i(S_j(t)u)\right)\Big). \end{equation} Consider the map \begin{equation}\label{starfunction2} (\bar\eta\star\hat\varphi)(t,x_{1},\ldots,x_{a+b}) = \big(\eta_{a+1}(t)\hat\varphi(t,x_{a+1}), \ldots, \eta_{a+b}(t)\hat\varphi(t,x_{a+b})\big)\in\mathbb{R}^{8b}, \end{equation} with $t\in I$ and $x\in \mathbb{R}^{n-1}=\mathbb{R}^{a+b}$. Since $T'_1(t)^2+T_2'(t)^2= \psi_1(t)^2+\psi_2(t)^2= 1$, by using Lemma \ref{lemmaderiv}, it follows similarly as in (\ref{etah}) that the pull-back symmetric tensor by the map $\bar\eta\star\hat\varphi:I\times \mathbb{R}^{n-1}\to \mathbb{R}^{8b}$ satisfies \begin{equation}\label{hatvp} (\bar\eta\star\hat\varphi)^*(\left\langle\,,\right\rangle) = 2\epsilon(t)^2\, dt^2 + \eta_{a+1}(t)^2dx_{a+1}^2+\ldots+\eta_{a+b}(t)^2dx_{a+b}^2, \end{equation} where $\epsilon:I\to [0,\infty)$ is the smooth function defined as in (\ref{ep_r}). We choose the step functions $S_1$ and $S_2$ so that (\ref{cond-ep}) is satisfied. This implies that $\rho(t)^2-2\epsilon(t)^2>0$. Let $\hat f:I\times \mathbb{R}^{n-1}\to \mathbb{R}^{8n-7-6a}_a=\mathbb{R}\times (\mathbb{R}^2_1)^a\times \mathbb{R}^{8b}$ be the map \begin{equation*} \hat f(t,x) = \left(\int_{t_0}^t\sqrt{\rho(\tau)^2+|\tilde\eta'(\tau)|^2-2\epsilon(\tau)^2}\,d\tau\,,\, \tilde\eta\star h\,(t,x)\,,\, \bar \eta\star \hat\varphi\,(t,x)\right), \end{equation*} where $\tilde\eta\star h:I\times \mathbb{R}^{n-1}\to (\mathbb{R}^2_1)^a$ is the map defined as in (\ref{starfunction}). If $b=0$, we define $\hat f(t,x)$ by simply omitting $\epsilon(t)$ and $\bar\eta\star\hat\varphi(t,x)$ in the definition of $\hat f(t,x)$ above. By using (\ref{etah}) and (\ref{hatvp}), it is easy to conclude that $\hat f:M^n\to \mathbb{R}^{8n-7-6a}_a$ is an isometric immersion. \begin{claim}\label{hat f inj} The immersion $\hat f$ is injective. \end{claim} In fact, assume that $\hat f(t^1,x^1)=\hat f(t^2,x^2)$, for some $t^1,t^2\in I$ and $x^1,x^2\in \mathbb{R}^{n-1}$. We write $x^j=(x^j_1,\ldots,x^j_{a+b})$, with $j=1,2$. Using that the function \begin{equation*} s(t)=\int_{t_0}^t\sqrt{\rho(\tau)^2+|\tilde\eta'(\tau)|^2-2\epsilon(\tau)^2}\,d\tau, \ t\in I, \end{equation*} is increasing, we obtain $t^1=t^2$. Since $\psi_1^2+\psi_2^2=1$, we can assume, without loss of generality, that $\psi_1(t^1)\neq 0$. Using that $\eta_i(t)>0$, for all $i=1,\ldots,a+b$ and $\hat f(t^1,x^1)=\hat f(t^1,x^2)$, we have $h(x^1_k)=h(x^2_k)$ and $\varphi_{11}(t^1,x^1_r)=\varphi_{11}(t^1,x^2_r)$, for all $k=1,\ldots,a$ and $r=a+1,\ldots,a+b$. These imply that \begin{equation*} \begin{array}{l} \sinh(x^1_k)=\sinh(x^2_k) \ \mbox{ and } \ \sin(T_1(S_1(t^1)x^1_r))=\sin(T_1(S_1(t^1)x^2_r)), \end{array} \end{equation*} for all $k=1,\ldots,a$ and $r=a+1,\ldots,a+b$. Since $S_1(t^1)>0$ and the functions $\sinh(u)$ and $\sin(T_1(u))$, with $u\in\mathbb{R}$, are injective, we obtain that $x^1=x^2$. Claim \ref{hat f inj} is proved. \begin{claim}\label{hat f emb} $\hat f:M^n\to \mathbb{R}^{8n-7-6a}_a$ is an isometric embedding. \end{claim} We just need to show that the inverse map $\hat f^{-1}:\hat f(I\times \mathbb{R}^{n-1})\to I\times \mathbb{R}^{n-1}$ is continuous. In fact, let $y_m=\hat f(t_m,x^m_1,\ldots,x^m_{n-1})$ be a sequence that converges to a point $y_\infty=\hat f(t_\infty,x^\infty_1,\ldots,x^\infty_{n-1})$. Since the function $s(t)$ is the first coordinate of $\hat f(t,x)$, we obtain $\lim s(t_m)=s(t_\infty)$. This implies that $\lim t_m=t_\infty$, since $s:I\to \mathbb{R}$ is a diffeomorphism of $I$ onto its image $s(I)$. Since the coordinates of the map $\eta(t)=(\eta_1(t),\ldots,\eta_{n-1}(t))$ are positive and smooth, we obtain \begin{enumerate}[(a)] \item\label{it-c} $\lim h(x^m_k)=h(x^\infty_k)$ \item\label{it-b} $\lim \varphi_{ji}(t_m,x^m_r)=\varphi_{ji}(t_\infty,x^\infty_r)$, \end{enumerate} for all $i,j=1,2$, \, $k=1,\ldots,a$\, and \, $r=a+1,\ldots,n-1$. It follows from \ref{it-c} that $\lim x^m_k=x^\infty_k$, for all $k=1,\ldots,a$, since $h(u)=(\cosh(u),\sinh(u))$ and $\sinh(u)$, with $u\in \mathbb{R}$, is a diffeomorphism. Now, using that $\psi_1(t_\infty)^2+\psi_2(t_\infty)^2=1$, we can assume that $\psi_1(t_\infty)\neq 0$. Since $\psi_1(t_\infty)>0$, we obtain that $S_1$ is a positive constant function in a neighborhood of $t_\infty$. This implies that $S_1(t_m)=S_1(t_\infty)>0$, for sufficiently large $m$. Since $\lim \psi_1(t_m)=\psi_1(t_\infty)>0$, we obtain from \ref{it-b} and (\ref{def-vp-jk}) that \begin{eqnarray*} \lim \cos(T_1(S_1(t_\infty)x^m_r))&=& \lim \frac{S_1(t_m)}{\psi_1(t_m)}P (\varphi_{11}(t_m,x^m_r)) = \frac{S_1(t_\infty)}{\psi_1(t_\infty)}P(\varphi_{11}(t_\infty,x^\infty_r)) \\&=& \cos(T_1(S_1(t_\infty)x^\infty_r)), \end{eqnarray*} for all $r=a+1,\ldots,n-1$, where $P:\mathbb{R}^2\to \mathbb{R}$ is the projection $P(u,v)=u$. Again using that $S_1(t_\infty)> 0$ and since $\cos(T_1(u))$ is a diffeomorphism of $\mathbb{R}$ onto $(0,1)$, it follows that $\lim x^m_r=x^\infty_r$, for all $r=a+1,\ldots,n-1$. We conclude that $\hat f^{-1}$ is continuous. Claim \ref{hat f emb} is proved. \\ Let $\hat f_h: I\times \mathbb{R}^{n-1}\to \mathbb{R}^{8n-6-5a}_{a+1}=\mathbb{R}^2_1\times \mathbb{R}^a\times (\mathbb{R}^2_1)^a\times \mathbb{R}^{8b}$ be the map \begin{equation}\label{def hat f h} \hat f_h(t,x) = \left(\sqrt{\frac{1}{c}+2\alpha(t)}\,h(\hat\theta_h(t))\,,\, \tilde\eta(t) \,,\,\, \tilde\eta\star h\,(t,x)\,,\, \bar \eta\star \hat\varphi\,(t,x)\right), \end{equation} where $\alpha:I\to [0,\infty)$ is as defined in (\ref{def-a}) and $\hat\theta_h:I\to \mathbb{R}$ is the function given by \begin{equation} \hat\theta_h(t)=\int_{t_0}^t \sqrt{\frac{1}{\frac{1}{c}+2\alpha(\tau)}\left(\rho(\tau)^2-2\epsilon(\tau)^2+\frac{\alpha'(\tau)^2}{\frac{1}{c}+2\alpha(\tau)}\right)}\,d\tau. \end{equation} If $b=0$, we define $\hat f_h(t,x)$ simply by omitting $\alpha(t)$, $\epsilon(t)$ and $\bar\eta\star\varphi(t,x)$ in the expressions of $\hat\theta_h(t)$ and $\hat f_h(t,x)$ above. By (\ref{cond-ep}), we have $\rho(t)^2-2\epsilon(t)^2>0$. This implies that $\hat\theta_h:I\to \mathbb{R}$ is well defined, smooth and increasing. Note that $\left\langle \hat f_h(t,x)\hat f_h(t,x)\right\rangle = - \left(\frac{1}{c}+2\alpha(t)\right) + |\tilde\eta(t)|^2 - |\tilde\eta(t)|^2 + 2\alpha(t) = -\frac{1}{c}$. Thus the image of $\hat f_h$ is contained in $\mathbb{H}^{8n-7-5a}_a(-c)$. By a standard computation, \begin{eqnarray*} (\hat f_h)^*(\left\langle\,,\right\rangle) &=& \left(-\frac{\alpha'(t)^2}{\frac{1}{c}+2\alpha(t)} + \left(\frac{1}{c}+2\alpha(t)\right)\hat\theta_h'(t)^2+|\tilde\eta'(t)|^2\right) dt^2 \\&& \,+\, (\tilde\eta\star h)^*(\left\langle\,,\right\rangle) + (\bar\eta\star \hat\varphi)^*(\left\langle\,,\right\rangle)\\&=& \rho(t)^2 dt^2 + \eta_1(t)^2 dx_1^2+\ldots+\eta_{a+b}(t)^2dx_{a+b}^2. \end{eqnarray*} Thus $\hat f_h:M^n\to \mathbb{H}^{8n-7-5a}_a(-c)$ is an isometric immersion. \begin{claim}\label{hat f h inj} The immersion $\hat f_h$ is injective. \end{claim} In fact, assume that $\hat f_h(t^1,x^1)=\hat f_h(t^2,x^2)$. Using (\ref{def hat f h}), we have \begin{equation*} \sqrt{\frac{1}{c}+2\alpha(t^1)}\,h(\hat\theta_h(t^1)) = \sqrt{\frac{1}{c}+2\alpha(t^2)}\,h(\hat\theta_h(t^2)). \end{equation*} Since $\left\langle h(u), h(u)\right\rangle=-1$, for all $u\in \mathbb{R}$, we obtain $\frac{1}{c}+2\alpha(t^1)=\frac{1}{c}+2\alpha(t^2)$, hence $h(\hat\theta_h(t^1))=h(\hat\theta_h(t^2))$. This implies that $t^1=t^2$, since the function $\sinh(\hat\theta_h(t))$, with $t\in I$, is increasing. The argument to show that $x^1=x^2$ is similar the one as given in Claim \ref{hat f inj}. Claim \ref{hat f h inj} is proved. \begin{claim}\label{hat f h emb} $\hat f_h:M^n\to \mathbb{H}^{8n-7-5a}_a(-c)$ is an isometric embedding. \end{claim} In fact, we just need to prove that the inverse map $(\hat f_h)^{-1}: \hat f_h(I\times \mathbb{R}^{n-1})\to I\times \mathbb{R}^{n-1}$ is continuous. Let $y_m=\hat f_h(t_m,x^m)$ be a sequence that converges to a point $y_\infty=\hat f_h(t_\infty,x^\infty)$. Using (\ref{def hat f h}), we obtain \begin{equation*} \lim \sqrt{\frac{1}{c}+2\alpha(t_m)}\, h(\hat\theta_h(t_m))=\sqrt{\frac{1}{c}+2\alpha(t_\infty)}\, h(\hat\theta_h(t_\infty)). \end{equation*} Using $\left\langle h(u),h(u)\right\rangle=-1$, we obtain $\lim (\frac{1}{c}+2\alpha(t_m))=\frac{1}{c}+2\alpha(t_\infty)>0$, hence $\lim h(\hat\theta_h(t_m))=h(\hat\theta_h(t_\infty))$. This implies that $\lim \sinh(T_1(\hat\theta_h(t_m)))=\sinh(T_1(\hat\theta_h(t_\infty)))$. Using that $\sinh(T_1(\hat\theta_h(t)))$ is a diffeomorphism of $I$ onto its image, it follows that $\lim t_m=t_\infty$. The argument to show that $\lim x^m=x^\infty$ is also similar to Claim \ref{hat f emb}. Thus Claim \ref{hat f h emb} is proved. \\ Now let $\hat f_s:I\times \mathbb{R}^{n-1}\to \mathbb{R}^{8n-4-6a}_a=\mathbb{R}^4\times (\mathbb{R}^2_1)^a\times \mathbb{R}^{8b}$ be the map \begin{equation}\label{def-hat f s} \hat f_s(t,x) = \left(\sqrt{\frac{1}{2c}+\frac{|\tilde\eta(t)|^2}{2}-\alpha(t)}\, \mathcal C(\hat\theta_s(t))\,,\,\,\tilde\eta\star h\,(t,x)\,,\,\bar\eta\star \hat\varphi\,(t,x)\right), \end{equation} where $\mathcal C(u)=(\cos(T_1(u)),\sin(T_1(u)), \cos(T_2(u)),\sin(T_2(u)))$, with $u\in \mathbb{R}$. Further, the function $\alpha:I\to [0,\infty)$ is given as in (\ref{def-a}), and $\hat\theta_s:I\to \mathbb{R}$ is defined by \begin{equation*} \hat\theta_s(t)=\int_{t_0}^t \sqrt{\frac{G(\tau)}{\frac{1}{2c}+\frac{|\tilde\eta(\tau)|^2}{2} - \alpha(\tau)} }\,d\tau, \end{equation*} where $G:I\to \mathbb{R}$ is the function $$G(t)=\rho(t)^2 +|\tilde\eta'(t)|^2 - 2\epsilon(t)^2 - 2\frac{\left(\left(-\alpha(t)+\frac{1}{2}|\tilde\eta(t)|^2\right)'\right)^2}{4(\frac{1}{2c}+\frac{|\tilde\eta(t)|^2}{2}-\alpha(t))}.$$ If $b=0$, we define $\hat f_s$ by simply omitting $\alpha(t)$ and $\bar\eta\star\hat\varphi(t,x)$ in the definitions of $\hat\theta_s(t)$ and $\hat f_s(t,x)$ above. We claim that we can choose the step functions $S_1$ and $S_2$ sufficiently large so that $\hat\theta_s$ is well defined and smooth. By (\ref{gransphere}), we already have $\frac{1}{2c}-\alpha(t)>0$. Furthermore, by a simple computation, \begin{eqnarray*} G(t) &=& \rho(t)^2 + |\tilde\eta'(t)|^2 - \frac{\left\langle\tilde\eta'(t),\tilde\eta(t)\right\rangle^2}{\frac{1}{c}+|\tilde\eta(t)|^2-2\alpha(t)} - \Delta(t)\\ &\ge & \rho(t)^2 + |\tilde\eta'(t)|^2 \left[1-\frac{|\tilde\eta(t)|^2}{\frac{1}{c}+|\tilde\eta(t)|^2-2\alpha(t)}\right] - \Delta(t),\end{eqnarray*} where $\Delta(t)=2\epsilon(t)^2 + \frac{\alpha'(t)^2 - 2\alpha'(t)\left\langle \tilde\eta'(t),\tilde\eta(t)\right\rangle}{\frac{1}{c}+|\tilde\eta(t)|^2-2\alpha(t)}$. Note that we can take $\epsilon(t)$, $\alpha(t)$ and $\alpha'(t)$ as smaller as we want if $S_1(t)$ and $S_2(t)$ become larger. Thus, we can choose the step functions $S_1$ and $S_2$ sufficiently large so that $\Delta(t)<\rho(t)^2$. This implies that $\hat \theta_s (t)$ is well defined, smooth and increasing. Note that $\left\langle \hat f_s(t,x),\hat f_s(t,x)\right\rangle = \frac{1}{c}$ since $|\mathcal C(u)|^2=2$, $|\bar\eta\star \hat\varphi\,(t,x)|^2=2\alpha(t)$ and $\left\langle \tilde\eta\star h\,(t,x)\,,\,\tilde\eta\star h\,(t,x) \right\rangle=-|\tilde\eta(t)|^2$. Thus the image $\hat f_s(I\times \mathbb{R}^{n-1})\subset \mathbb{S}^{8n-5-6a}_a(c)$. By a direct computation, we show that \begin{eqnarray*} \hat f_s^*(\left\langle\,,\right\rangle) &=& \left(\frac{\left(\left(-\alpha(t)+\frac{1}{2}|\tilde\eta(t)|^2\right)'\right)^2}{\frac{1}{c}+|\tilde\eta(t)|^2-2\alpha(t)} + \left(\frac{1}{2c}+\frac{|\tilde\eta(t)|^2}{2}-\alpha(t)\right)\hat\theta_s'(t)^2\right)dt^2\\&& + (\tilde\eta\star h)^*(\left\langle\,,\right\rangle) + (\bar\eta\star \hat\varphi)^*(\left\langle\,,\right\rangle)\\&=& \rho(t)^2 dt^2 + \eta_1(t)^2 dx_1^2 +\ldots+\eta_{a+b}(t)^2 dx_{a+b}^2. \end{eqnarray*} This implies that $\hat f_s:M^n\to \mathbb{S}^{8n-5-6a}_a(c)$ is an isometric immersion. \begin{claim}\label{hat f s inj} The immersion $\hat f_s$ is injective. \end{claim} In fact, assume that $\hat f_s(t^1,x^1)=\hat f_s(t^2,x^2)$, for some $t^1,t^2\in I$ and $x^1,x^2\in \mathbb{R}^{n-1}$. Using (\ref{def-hat f s}), $$\sqrt{\frac{1}{c}+|\tilde\eta(t^1)|^2-2\alpha(t^1)}\, \mathcal C(\hat\theta_s(t^1))=\sqrt{\frac{1}{c}+|\tilde\eta(t^2)|^2-2\alpha(t^2)}\, \mathcal C(\hat\theta_s(t^2)).$$ Since $|\mathcal C(\hat\theta_s(t))|^2=2$, for all $t\in I$, we obtain $|\tilde\eta(t^1)|^2-2\alpha(t^1)=|\tilde\eta(t^2)|^2-2\alpha(t^2)$, hence $\mathcal C(\hat\theta_s(t^1))=\mathcal C(\hat\theta_s(t^2))$. This implies that $t^1=t^2$, since $\sin(T_1(\hat\theta_s(t)))$, with $t\in I$, is increasing. The argument to show that $x^1=x^2$ is similar to that one given in Claim \ref{hat f inj}. \\ \begin{claim}\label{hat f s emb} $\hat f_s:M^n\to \mathbb{S}^{8n-5-6a}_a(c)$ is an isometric embedding. \end{claim} In fact, we just need to prove that the inverse map $(\hat f_s)^{-1}: \hat f_s(I\times \mathbb{R}^{n-1})\to I\times \mathbb{R}^{n-1}$ is continuous. Let $y_m=\hat f_s(t_m,x^m)$ be a sequence that converges to a point $y_\infty=f(t_\infty,x^\infty)$. Using (\ref{def-hat f s}), $$\lim \sqrt{\frac{1}{c}+|\tilde\eta(t_m)|^2-2\alpha(t_m)}\, \mathcal C(\hat\theta_s(t_m))=\sqrt{\frac{1}{c}+|\tilde\eta(t_\infty)|^2-2\alpha(t_\infty)}\, \mathcal C(\hat\theta_s(t_\infty)).$$ Since $|\mathcal C(\hat\theta_s(t))|^2=2$, for all $t\in I$, we have $\lim (|\tilde\eta(t_m)|^2- 2\alpha(t_m))=|\tilde\eta(t_\infty)|^2- 2\alpha(t_\infty)>0$, hence $\lim \mathcal C(\hat\theta_s(t_m))=\mathcal C(\hat\theta_s(t_\infty))$. This implies that $\lim \sin(T_1(\hat\theta_s(t_m)))=\sin(T_1(\hat\theta_s(t_\infty)))$. Using that $\sin(T_1(\hat\theta_s(t)))$ is a diffeomorphism of $I$ onto its image, it follows that $\lim t_m=t_\infty$. The argument to show that $\lim x^m=x^\infty$ is similar to Claim \ref{hat f emb}. Thus, Claim \ref{hat f s emb} is proved. Theorem \ref{mth} is proved.
\section{Introduction} Almost all of the many variants of projector Quantum Monte Carlo (QMC) rely on the properties of the operator $e^{-\beta H}$, where due to its similarity to the time-evolution operator, the variable $\beta$ is denoted `imaginary time'. Generally, this imaginary time is discretized, and the operator iteratively applied as a short-time propagator, in order to simulate its action in the large $\beta$ limit\cite{Foulkes2001}. Expressing an initial wavefunction in the eigenbasis of the Hamiltonian of interest, the application of this $e^{-\beta H}$ propagator results in a projection onto the $i^{\textrm{th}}$ eigenstate proportional to $e^{-\beta E_i}$, where $E_i$ is the energy of this eigenstate. It is clear to see that in this large $\beta$ limit, the projection onto the lowest energy eigenvector dominates the wavefunction, whereby ground state properties can be extracted, assuming some overlap with the initial wavefunction. While this formalism is clearly powerful, by construction it exponentially quickly projects out excited states of the system which may be of interest, and are of critical importance in the simulation of finite-temperature properties, reaction dynamics, photochemistry and many other areas. To date, isolating excited states of systems via projector QMC methods has only been practical with a restriction on the projection to sample a space which is approximately orthogonal to those of the lower energy states, via nodal constraint\cite{Lester1986}, or orthogonalization against them in a subspace projection method\cite{Ceperley1988,Nagase2010}. More indirectly, statistical methods have been used on short periods of imaginary-time in order to isolate individual decay rates in the spectrum by analytic continuation to a real-time dynamic\cite{Baym1961,Gubernatis91}. However, these approaches are not entirely satisfactory; accurate nodal surfaces for excited states can be difficult to obtain, resulting in a larger fixed node error and potentially transient estimators, while the subspace projection method has limited applicability\cite{Ceperley1997}. In addition, the Bayesian techniques which rely on maximizing entropy to approximate a notoriously unstable inverse Laplace transform, have difficulty achieving quantitative accuracy within noisy data sets \cite{Berne1983,Blume1997,Gunnarsson10}. Despite this, there are examples of accurate excited states within the nodal constraint\cite{Umrigar2009}, while maximum entropy techniques are particularly prevalent in solid state calculations to obtain the density of states, often in the case of continuous-time QMC as applied to quantum impurity models and dynamical mean-field theory\cite{Liu2012,Millis2011}. Here, we take a different approach to the calculation of excited states, within the context of Full Configuration Interaction Quantum Monte Carlo (FCIQMC)\cite{BTA2009,CBA2010,BoothC2}. This recently introduced method applies the imaginary-time evolution propagator to a stochastic `walker' representation of the wavefunction expressed in the full space of Slater determinants constructed from a single-particle basis of size $M$. Although this reintroduces a basis set error compared to those methods operating in real space, it confers various advantages which mitigate this. The discrete basis allows for an efficient walker annihilation algorithm, which can exactly overcome the fermion sign problem in the sampling, provided enough walkers are used\cite{Spencer2012}. The `initiator' approximation was formulated to maintain a high annihilation rate, and control the growth of noise in a systematically improvable fashion\cite{CBA2010,BoothC2}. This has allowed far larger systems to be treated at an accuracy comparable to that of Full Configuration Interaction (FCI or exact diagonalization), within small and systematically improvable error bars. In addition, a semi-stochastic adaptation of the algorithm\cite{UmrigarArXiv}, as well the introduction of a partial nodal constraint\cite{Clark2012} and ideas from quantum chemistry\cite{BoothF12} hold promise of increased accuracy and efficiency of the method. In order to project out a targeted excited state rather than the ground state, we propose the use of a projection operator of the form \begin{equation} P(H)=e^{-\beta^2(H-S)^2} . \label{eqn:Propagator} \end{equation} For sufficiently large $\beta$, this Gaussian propagator will result in the dominant eigenstate being the one closest in energy to the chosen value of the diagonal offset $S$, termed the shift. In the eigenbasis of $H$, $\{ |\Psi_i\rangle, E_i \}$ with $\Psi_0$ representing the ground state, and starting from an initial wavefunction ${ |\psi_{\textrm{T}} \rangle }$ with $S=E_k$, it can be seen that the long time propagation results in \begin{equation} |\Psi_k\rangle \propto \lim_{\beta\rightarrow\infty} \sum_i |\Psi_i\rangle e^{-\beta^2(E_i-E_k)^2} \langle \Psi_i { |\psi_{\textrm{T}} \rangle } \propto \sum_i \delta_{E_i,E_k} |\Psi_i \rangle . \end{equation} We note here that a projector of this form was proposed back in 1983 within continuum QMC approaches\cite{Hirsch1983,Berne1985}, although due to sign problems, and significantly larger timestep errors resulting from the fact that $H^2$ is more singular than $H$, only one-electron systems were reported, and no modern implementation exists in the literature. This issue of the timestep highlights another advantage of working in a finite basis, in that the spectrum is bounded both from below and above. This allows for linearization of the short-time propagator, \begin{eqnarray} |\Psi\rangle &\propto& \lim_{P\rightarrow\infty} \left[A e^{-\tau (H-S)^2}\right]^P { |\psi_{\textrm{T}} \rangle } \\ |\Psi\rangle &\propto& \lim_{P\rightarrow\infty} \left[A(1-(\tau (H-S))^2)\right]^P { |\psi_{\textrm{T}} \rangle } , \end{eqnarray} without becoming unbound and dominated by very high energy states oscillating in time, and without incurring timestep errors in the final wavefunction so long as the timestep is less than an upper bound given by $\tau \leq \frac{2}{E_{\mathrm{max}}-E_{\mathrm{min}}}$. Repeated application of the short-time propagator therefore results in a `power-method' for states on the interior of the spectrum, rather than at the extrema, with similarities to filter diagonalization\cite{Grosso1993,Neuhauser95}. Propagation with \rff{eqn:Propagator} leads to a theoretical decay of state $j$ from $i$ as $e^{-\beta^2((E_i-S)^2-(E_j-S)^2)}$. In contrast with the ground state propagator, this rate of decay depends on $S$, with it being advantageous to choose $S$ to be as close to the energy of the state of interest as possible. However, even if $S$ is chosen exactly, the projection of the non-dominant states is slower compared to the ground state propagation, and we will return to this issue later. In addition, unless $S$ is chosen exactly, the long-time propagation of the dynamic will result in a continued projection onto a decaying function of all states, including the dominant one. For this reason, the factor of $A$ is introduced into the short-time propagator, such that at convergence, its value can be varied in order to maintain a constant L$^1$ normalization of the dominant wavefunction and a stable number of walkers. This is analogous to the variation of $S$ in the ground state projection\cite{BTA2009}. The differential formulation of the exact dynamic governed by this propagator for a given component of the wavefunction, $C_i$, can be written as \begin{equation} \frac{d C_i}{d \tau^2} = (A-1)C_i - \epsilon A \sum_{j,k} (H_{ij}-\delta_{ij}S)(H_{jk}-\delta_{jk}S) C_k , \label{eqn:Dynamic} \end{equation} where $\epsilon \rightarrow \tau^2$ as $A \rightarrow 1$, and the application of $H^2$ has been decomposed by a resolution of the identity over the connecting space of determinants $j$. This formulation is now amenable to stochastic integration with a discrete walker representation of the determinantal wavefunction coefficients $C$. As with the ground state projection, there is no unique stochastic algorithm for this dynamic, but the one which we found to be most efficient involves a double spawning cycle, which requires little additional overhead compared to the ground state algorithm, and no additional memory requirements. Each iteration, the determinants represented by $k$ are run through, and a spawning step attempted to determinant $j$, in the same fashion as the ground state propagation, but in negative time. This results in a spawning probability to a connected determinant $j$ with a stochastically realised {\em signed} amplitude of $\frac{\tau H_{jk}}{P(k|j)}$, where $P(k|j)$ represents the normalised probability to randomly select symmetry-connected determinant $j$ from $i$. Successfully spawned walkers are subsequently propagated again in the same iteration with a now forwards-time signed amplitude of $-\frac{\tau H_{ji}}{P(j|i)}$. Care must be taken that for determinant $k$, the diagonal `death' processes from the first application of $H$ are now interpreted as spawning events, which are also subsequently propagated via $H_{ij}$. Each iteration, the factor of $A$ is applied initially as a separate enhancement of the local population of each determinant, with the absolute population on each determinant growing with probability $A C_k$. \begin{figure}[t] \begin{center} \includegraphics[scale=0.475]{plot-It_Eigenfuncoverlap.eps} \end{center} \caption{Convergence of the propagation to the second excited state of He$_2$ at 2.5~\AA\ separation in a cc-pVDZ basis. $S$ was fixed at -3.65${\textrm{E}_{\textrm{h}}}$, and $A$ at 1.004 until 10,000 walkers were present, denoted by the dotted line, where $A$ was varied in order to keep this number constant. After variation, the average value of $A-1$ was 1.4(6)$\times10^{-6}$. } \label{He2conv} \end{figure} In Fig.~\ref{He2conv}, we present an illustrative example of the algorithm for the helium dimer in a cc-pVDZ basis, small enough such that the full spectrum of eigenstates can be calculated and the convergence of the method analysed. The value of $S$ was fixed at $\sim40{\textrm{mE}_{\textrm{h}}}$ higher than the second excited state, but such that it remained the dominant state in the dynamic, and was subsequently found to be correctly projected out over time. This is despite working in a canonical representation, and starting with a single walker on the Hartree--Fock determinant which had an initial overlap with the ground state of close to one. In order to grow the walkers, $A$ was initially fixed at a value of 1.004, and was varied when a target number of walkers was reached, in common with the procedure for the ground state propagation. The convergence of the projected energy, as defined in Ref.~\onlinecite{BTA2009}, is shown in Fig.~\ref{He2Energy} for the same simulation, and reflects the decay from the sampled wavefunction of the first excited state. \begin{figure}[t] \begin{center} \includegraphics[scale=0.475]{plot-EnergyConvergeHF.eps} \end{center} \caption{ Convergence of the projected energy estimate to the exact eigenvalue. The reference determinant for the projection was dynamically adjusted to project onto the largest weighted determinant in the sample. } \label{He2Energy} \end{figure} In order to reliably extend to larger molecular systems, it is worth considering how to transfer the salient elements of the initiator approximation into this new propagation. The basis of this approximation is to attempt to propagate walkers corresponding to wavefunction signal exactly, while walkers judged to be potentially noise are propagated with a truncated Hamiltonian which acts only over the instantaneously occupied subspace\cite{CBA2010,BoothC2}. This is systematically improvable as the instantaneously occupied subspace grows, or the criterion for walkers corresponding to signal becomes more inclusive. The separation between walkers corresponding to signal and potential noise is not unique. However, it seems sensible to retain the tested feature from ground state propagation that newly spawned walkers on previously unoccupied determinants ($i$) at the end of an iteration, must have come from an initial determinant ($k$) which is deemed to have a well-established sign, and therefore a population of walkers above ${n_{\textrm{add}}}$. Consequently, all walkers from the application of the first Hamiltonian operator are kept, while the information as to whether $C_i$ is above the initiator criterion is passed through to the annihilation stage of the final set of spawned walkers. No walkers are therefore aborted over the resolution of the identity between $H^2$ (determinants $j$ in \rff{eqn:Dynamic}). To test this on a larger system, we study an excited state deep in the spectrum of the 10-electron neon atom in a cc-pVDZ basis, with an energy of approximately $2.5{\textrm{E}_{\textrm{h}}}$ above the ground state. Setting $S$ to equal the CISDTQ energy for the corresponding state, and while remaining in a canonical Hartree--Fock basis and starting from a random distribution of walkers throughout the whole space, we achieved a converged energy of -126.2118(4)${\textrm{E}_{\textrm{h}}}$, compared to the FCI value of -126.21177${\textrm{E}_{\textrm{h}}}$. This value is $4.76{\textrm{mE}_{\textrm{h}}}$ lower than the initial guess provided by CISDTQ. It would be highly advantageous to develop a robust algorithm for varying $S$ dynamically during the run, as is done for the ground state algorithm. This could be used to maximise the rate of growth of walkers or alternatively minimize $A$, both of which should adjust $S$ to more closely match the eigenvalue of the state, remove reliance on the initial guess and increase the convergence rate. However, since this requires finding a minimum in a quadratically varying and noisy dataset, no robust algorithm has been identified so far. Dynamic correlation and response functions due to some perturbation, either in the frequency or time domain, are of critical importance in electronic structure theory\cite{Balseiro1987}, and are directly measured in experiments to probe the electronic properties of materials through techniques such as neutron scattering or photoelectron spectroscopy\cite{Hutchings1972}. Many methods, including in general QMC approaches, have significant difficulty in calculating these quantities\cite{Berne1991}, often having to rely on unstable analytic continuation from imaginary time correlation functions\cite{Baym1961,Berne1983,Gubernatis91,Gunnarsson10,Blume1997}, while other methods can bias towards high or low energy regimes\cite{Millis2011}. Although other correlation functions are possible, here we look at the example of an advanced Green's function, a central concept in electronic structure where the `perturbation' at time $t=0$ is the creation of a hole in orbital $j$. For negative time periods, $t$, these can be written in the time and frequency domain respectively as \begin{eqnarray} G^-(i,j,t) &=& i\langle \Psi_0|a^{\dagger}_i e^{-i(H-E_0-i\delta)t} a_j| \Psi_0 \rangle \\ G^-(i,j,\omega) &=& \langle \Psi_0|a^{\dagger}_i \frac{1}{\omega-(H-E_0)+i\delta} a_j| \Psi_0 \rangle . \label{eqn:FreqGF} \end{eqnarray} Unlike the inverse Laplace transform required for the analytic continuation of imaginary time correlation functions, the transform between these two domains is a well-conditioned and numerically stable fourier transform in the presence of noisy data. Spectral density functions, such as the density of states for extended systems, are then defined in the Lehmann representation\cite{FetterandW} as \begin{eqnarray} A^-(i,j,\omega) &=& -\frac{1}{\pi} \Im[G^-(i,j,\omega)] \\ &=& \frac{1}{\pi}\sum_{n}\frac{\langle \Psi_0^{N}|a^{\dagger}_i|\Psi_n^{N-1}\rangle \delta \langle \Psi_n^{N-1} | a_j | \Psi_0^{N} \rangle}{(\omega-E_n^{N-1}+E_0^N)^2+\delta^2} , \label{eqn:ShortTimeSpectral} \end{eqnarray} which in the small $\delta$ limit tends to \begin{align} A^-(i,j,\omega) =& \sum_{n}\langle \Psi_0^{N}|a^{\dagger}_i|\Psi_n^{N-1}\rangle \langle \Psi_n^{N-1} | a_j | \Psi_0^{N} \rangle \times \notag \\ &\delta(\omega-(E_n^{N-1}-E_0^N)) , \label{eqn:Spectral} \end{align} where $\delta$ in the above equation represents the dirac-delta function. Assuming $A=1$, application of the propagator in \rff{eqn:Propagator} for a time $\beta^2=\frac{1}{2 \delta^2}$ will result in the wavefunction \begin{equation} C(\beta^2)=e^{-\frac{1}{2 \delta^2}(H-S)^2}{ |\psi_{\textrm{T}} \rangle } , \label{eqn:IntDynamic} \end{equation} which when applied to an initial wavefunction ${ |\psi_{\textrm{T}} \rangle }=a_j|\Psi_0 \rangle$ obtained from the ground-state dynamic, and then projected onto $\frac{\beta}{\sqrt{\pi}}\langle \Psi_0 |a^{\dagger}_i$ will result in the distribution \begin{align} f(i,j,\omega) =& \frac{1}{\sqrt{2 \pi} \delta} \sum_{n}\langle \Psi_0^{N}|a^{\dagger}_i|\Psi_n^{N-1}\rangle \langle \Psi_n^{N-1} | a_j | \Psi_0^{N} \rangle \times \notag \\ & e^{-\frac{1}{2 \delta^2}(\omega-E_n^{N-1}+E_0^N)^2} \label{eqn:GF} \end{align} for $S=\omega+E_0^N$. This will tend to the spectral function given in \rff{eqn:Spectral} in the large $\beta$ limit. The real parts of the Green's function can then by obtained if needed from the Kramers-Kronig relation\cite{Balseiro1987}. We note that a related Green's function can be obtained directly by integrating \rff{eqn:IntDynamic} over $\beta$, with the addition of the small imaginary component $\delta$ to the dynamic. Unfortunately however, this integral is only convergent for $(\omega-E_n^{N-1}+E_0^N) > \delta$, and so the FCIQMC calculation will blow up at the poles. In systems with a continuous spectra, this would not be appropriate, and so we do not pursue this approach here. Results from a pilot investigation of the beryllium dimer in a cc-pVDZ basis, where the exact Green's function can be obtained from complete diagonalization, are shown in Fig.~\ref{Beryllium}. \begin{figure}[t] \begin{center} \includegraphics[scale=0.475]{Be2GF.eps} \end{center} \caption{ High energy window of the spectral function $A^-(1,1,\omega)$ for exact propagation with $\delta=0.0141{\textrm{E}_{\textrm{h}}}$, and stochastic evaluation via FCIQMC for an equivalent time $\beta=50$a.u. for frozen-core Be$_2$ in a cc-pVDZ basis at 2.5\AA\ . Vertical lines indicate the difference between the ground state energy and the eigenvalues of the N-1 system symmetry connected in $G^-(1,1,\omega)$, although some are coupled too weakly to contribute significantly to the spectral function. Approximately 10 independent calculations at each value of $\omega$ were averaged to obtain the errorbars. } \label{Beryllium} \end{figure} In order to reduce the statistical error, it may be necessary to average over a small number of independent calculations at each frequency, and this can be combined with an elimination of the bias derived from choosing a correlated sample of $\langle \Psi_0|a^{\dagger}_i$ and $a_j| \Psi_0 \rangle$ \cite{Ceperley1988}, by taking the $\Psi_0$ samples on each side of \rff{eqn:GF} from different snapshots in imaginary time. In addition, by storing multiple wavefunctions of the type $\langle \Psi_0|a^{\dagger}_i$ at the same time, all $M^2$ single-particle Green's functions can be calculated at a cost of $\mathcal{O}[M]$ FCIQMC calculations per frequency point, without the expectation of any variation in accuracy between high and low energy regimes. However, despite modest successes, it is clear that obtaining converged results through the use of this operator is substantially more difficult than with the ground state projection. This is mainly due to an additional factor of $(\tau \Delta E)^{-1}$ in the number of iterations required to project out undesired states with energy gap $\Delta E$ for comparable accuracy to the ground state propagation. The result is that while in the ground state propagation excited states were filtered relatively quickly with only isolated convergence issues in the case of near degeneracy\cite{BoothC2}, the number of iterations required for excited state propagation are substantially increased, as well as the dynamic being less well-conditioned with respect to walker fluctuations. This is also exacerbated by a generally more multiconfigurational wavefunction which increases random error in the projected energy estimator\cite{BTA2009}. A more judicious choice of orbital basis and initial conditions optimized for the state of interest, as well as a multireference projected energy formulation\cite{UmrigarArXiv} would ameliorate many of these issues. In addition, there is the possibility of preconditioning techniques familiar from iterative diagonalization methods\cite{Davidson1975} being transferred into the stochastic dynamic, as well other operators, such as $e^{-\beta |H|}$, which should behave more efficiently and allow for extension to larger systems. Research in these directions is currently under way. It is clear that alternative propagators within the FCIQMC dynamic holds promise for obtaining accurate excited states. \section*{Acknowledgements}
\section{Introduction} Thermodynamics \cite{Callen} has a reputation as being a science complete in its basic laws. Less well known is thermodynamic fluctuation theory which follows from thermodynamics \cite{Landau}. This fluctuation theory may be represented by a unique thermodynamic information Riemannian metric \cite{Ruppeiner1995}, resulting in an invariant Riemannian thermodynamic curvature scalar $R$. Physically, $R$ reveals information about interatomic interactions, and I illustrate here with recent calculations of $R$ in pure fluids with thermodynamic data from the NIST fluid database \cite{Ruppeiner2012}. As was shown by Bekenstein \cite{Bekenstein}, Hawking \cite{Hawking}, and others, thermodynamics may be extended to black holes. Logically, thermodynamic fluctuation theory, including the physical interpretation of $R$ developed in fluid and spin systems, should extend likewise. This paper includes a tabulation of calculations of $R$ from general relativistic black hole solutions, for which no underlying microscopic models are known. I also discuss various claims of ``inconsistencies'' concerning divergences of black hole heat capacities, and cases with $R=0$. \section{Pure fluids} The basic structure in thermodynamic fluctuation theory in pure fluids is an infinite environment in some reference state ``0'', characterized by two independent intensive thermodynamic variables, and containing an open subsystem with constant volume $V$. This subsystem exchanges energy and particles with its environment, and its thermodynamic state fluctuates about an equilibrium characterized by maximum entropy. The probability of a fluctuation away from equilibrium is given by Einstein's famous formula \cite{Landau}: $\mbox{probability}\propto\mbox{exp}\left[-V(\Delta\ell)^2 /2 \right]$, where $(\Delta\ell)^2$ is the invariant, positive definite, thermodynamic information metric. Expressed in the pair of independent thermodynamic variables $x^1$ and $x^2$: $(\Delta \ell)^2 = g_{\mu\nu}\Delta x^{\mu}\Delta x^{\nu}$, where $\Delta x^\alpha\equiv (x^\alpha -x^\alpha_0)$ denotes the difference between the thermodynamic variables $x^{\alpha}$ of the subsystem and their equilibrium values $x^{\alpha}_0$. Denote by $S$, $X^1$, and $X^2$ the entropy, internal energy, and particle number, respectively, of the subsystem. $X^1$ and $X^2$ are each conserved quantities, and $S$ is additive between the subsystem and its environment. If $x^{\alpha}=X^{\alpha}$, then: \begin{equation} g_{\alpha\beta} = -\frac{1}{k_B V}\frac{\partial^2 S} {\partial X^{\alpha}\partial X^{\beta}},\label{30}\end{equation} \noindent where $k_B$ is Boltzmann's constant \cite{Ruppeiner1995}. $g_{\alpha\beta}$ transforms as a second-rank tensor. The thermodynamic metric Eq. (\ref{30}) leads to an invariant Riemannian thermodynamic curvature scalar $R$. $R=0$ for the ideal gas \cite{Ruppeiner1979}, and $|R|$ diverges as the correlation volume $\xi^3$ near the critical point \cite{Ruppeiner1995,Ruppeiner1979,Johnston2003}. $R$ is thus a measure of effective interatomic interactions. Model calculations reveal that the sign of $R$ corresponds to whether interactions are effectively attractive ($R<0$) or repulsive ($R>0$); see the tabulation in \cite{Ruppeiner2010}. The connection of $|R|$ to fluctuating structure size has also been established directly by means of a covariant thermodynamic fluctuation theory \cite{Ruppeiner1995}. Most recently \cite{Ruppeiner2012}, $R$ has been calculated in pure fluids using data based on experiment. Pure fluids have interatomic interaction potentials of the Lennard-Jones type, strongly repulsive at short range, attractive at long range, and with a minimum at a distance where the atoms in the condensed liquid and solid phases reside. A fluid typically has density smaller than that of a condensed state, and we thus expect an average attractive interatomic interaction, and negative $R$. Negative $R$ is indeed the norm in pure fluids. Figure 1 shows $R$ for hydrogen along the coexistence curve, with: 1) $R$ in both phases diverging to $-\infty$ at the critical point, 2) $|R|$ in both phases decreasing as the triple point is approached, with liquid phase values approaching the order of a molecular volume, and 3) a region of small positive $R$ in the condensed liquid phase near the triple point as the liquid organizes into a solid-like structure. One may now pose a puzzle: which atomic arrangements produce the thermodynamic curvatures seen in Fig. 1? Figure 2 suggests a solution encompassing most fluid situations: Fig. 2a shows widely spaced atoms where attractive interactions form loose fluctuating clusters of volume $|R|\propto\xi^3$. Such structures are typical of the critical point regime. Fig. 2b shows a tight cluster of atoms in the vapor phase prevented from collapse by repulsive interactions. Such clusters are present in water, but not in hydrogen \cite{Ruppeiner2012}. Fig. 2c shows a haphazard arrangement of atoms in a condensed liquid state. Fig. 2d shows an organized solid-like state. The liquid phase in Fig. 1 shows a change in sign of $R$ corresponding to a transition Fig. 2c $\to$ Fig. 2d. Results for $R$ in pure fluids are thus associated with simple microscopic patterns. Black hole physics does not have as well developed a theoretical microscopic structure. However, the calculation of $R$, and insights from fluid or spin models, might lend some guidance in this difficult domain. \begin{figure}[h] \begin{minipage}{18pc} \includegraphics[width=18pc]{Hydrogen.jpg} \caption{\label{label}$R$ for hydrogen along the coexistence curve in the liquid and vapor phases from the triple point to the critical point, with temperatures $T=T_t$ and $T=T_c$.} \end{minipage} \hspace{0.4cm} \begin{minipage}{18pc} \includegraphics[width=18pc]{RSketches.jpg} \caption{\label{label} a) Atoms pulled together by attractive interactions; b) atom cluster held up by repulsive interactions; c) disorganized liquid phase; and d) organized solid phase.} \end{minipage}\hspace{0pc} \end{figure} \section{Black holes} Black hole solutions to general relativity have well-defined thermodynamic properties. The role of the conserved thermodynamic variables is played by mass, angular momentum, and charge $(M, J, Q)$, respectively. Logically, black hole thermodynamics leads to thermodynamic fluctuation theory and an information metric Eq. (\ref{30}) (set $V=1$, and use $(M, J, Q)$ for the $X^\alpha$'s \cite{Ruppeiner2008}). This information metric produces the black hole thermodynamic curvature $R$ \cite{Ferrara1997,Aman2003}. A tentative attempt to physically interpret $R$ for black holes has been made in terms of the spatial size of correlated fluctuations on the event horizon \cite{Ruppeiner2008}. Perhaps the puzzle solving tried above in pure fluids can eventually lead to insight about possible black hole microstructures. Debated in black hole thermodynamics has been the possibility raised by Davies \cite{Davies1977} that the curve of diverging heat capacity $C_{J,\,Q}\equiv T(\partial S/\partial T)_{J,\,Q}$ in the Kerr-Newman black hole corresponds to a phase transition. Since diverging heat capacities are a feature of second-order phase transitions in fluid and spin systems, Davies' association would appear logical. However, in ordinary thermodynamics the connection between thermodynamic anomalies and phase transitions gets support from microscopic models. Without such models there can be no assurance of any phase transition. For black holes, if some heat capacity diverges, how could we be sure that we have not just made an inappropriate choice of thermodynamic variables, which reveals infinities with no really fundamental significance? What would we make of curves where one heat capacity diverges, but the other heat capacities stay regular? What if each heat capacity diverges along its own curve, as happens in the Kerr-Newman black hole \cite{Ruppeiner2008}? Which curve corresponds to a true phase transition? These questions identify difficulties associating diverging heat capacities with phase transitions. It seems simpler to look at diverging $|R|$ to indicate phase transitions, since $R$ has a unique status in identifying microscopic order, which in fluid and spin systems goes to the heart of the issue of phase transitions. Black hole solutions with $R$ identically zero have also given rise to debate; see \cite{Medved2008} for a review. Examples of such solutions are the BTZ and the Reissner-Nordstr$\ddot{\mbox{o}}$m \cite{Aman2003} black holes. If $R$ measures in some sense the range of interactions, then one might expect $|R|$ to be large for black hole thermodynamics, reflecting the gravitational forces present in these objects. But such reasoning need not obtain. In a classical black hole, the gravitating particles have collapsed to a central singularity, shrinking the interactions between them to zero volume. The statistics underlying the thermodynamics might reside on the event horizon, where unknown constituents might interact with each other by forces perhaps not gravitational. In this scenario, gravity might merely be a nonstatistical force holding the assembly together, and a result $R = 0$, where the unknown constituents move independently of each other, might not be so unreasonable. \begin{table} \centering \caption{\label{book} Properties of $R$ for black holes solutions from general relativity.} \begin{tabular}{@{}l*{15}{c}} \br Name of solution & Dimension & Variables & Stable & Sign of $R$ & $R=0$ & $ \frac{|R|\to\infty}{T \ne 0}$ \\ \mr BTZ \cite{Aman2003} & $2 + 1$ & ($M$, $J$) & yes & $0$ & - & no \\ Kerr \cite{Aman2003} & $3 + 1$ & ($M$, $J$) & no & $+$ & no & no \\ Reissner-Nordstr$\ddot{\mbox{o}}$m\cite{Aman2003} & $3 + 1$ & ($M$, $Q$) & no & $0$ & - & no \\ Kerr-Newman \cite{Ruppeiner2008,Mirza2007} & $3 + 1$ & ($M$, $J$, $Q$) & no & $+$ & no & no \\ K-AdS \cite{Sahay2010b} & $3 + 1$ & ($M$, $J$) & yes & $-$ & no & yes \\ RN-AdS \cite{Aman2003} & $3 + 1$ & ($M$, $Q$) & yes & $\pm$ & yes & yes \\ KN-AdS \cite{Sahay2010} & $3 + 1$ & ($M$, $J$, $Q$) & yes & $\pm$ & yes & yes \\ Black Hole \cite{Arcioni2005} & $4 + 1$ & ($M$, $J$) & no & $+$ & no & no \\ Small Black Ring \cite{Arcioni2005} & $4 + 1$ & ($M$, $J$) & no & $\pm$ & yes & yes \\ Large Black Ring \cite{Arcioni2005} & $4 + 1$ & ($M$, $J$) & yes & $-$ & no & yes \\ \br \end{tabular} \end{table} Table 1 summarizes a number of black hole calculations of $R$. Tabulated are the spatial ($+$ time) dimensions, whether or not there are regimes of thermodynamic stability (positive definite metric matrix Eq. (\ref{30})), the sign of $R$ (or an indication ``0'' if $R$ is identically zero), whether or not there are places where the sign of $R$ changes through zero, which might indicate a Hawking-Page phase transition \cite{Sahay2010}, and whether or not there are divergences $|R|\to\infty$ with $T\ne 0$. Most solutions here with $R$ not identically zero diverge to $\pm\infty$ at the extremal curve $T\to 0$. I omit string theory black hole solutions for simplicity. Table 1 shows a progression from the elegant BTZ black hole, stable for all states, and with identically zero $R$, to the older solutions Kerr, Reissner-Nordstr$\ddot{\mbox{o}}$m, and Kerr-Newman, none of which rise to the level of stability. (Omitted are cases of the Kerr-Newman black hole in which one of the parameters is held fixed at value not zero, while the others fluctuate; such do show interesting instances of stability \cite{Ruppeiner2008}). A richer structure, including stability, $R$ zero crossing, and $R$ divergences, clearly results from either increasing the number of dimensions, adding an asymptotic AdS scenario, and/or more complex topology. Too intricate to tabulate here are results for Myers-Perry black holes \cite{Aman2010}. \ack I thank George Skestos for travel support, and Bhupendra Tiwari for productive discussions. \section*{References}
\section{Introduction} {\bf Overview. } In this work, we give a very simple local search algorithm for ordering constraints satisfaction problems that works better than the random assignment for those instances of the ordering $k$-CSP problem, where each variable is used only a bounded number of times. To motivate the study of the problem, we first overview some known results for regular constraint satisfaction problems. An instance of a constraint satisfaction problem consists of a set of variables $V = \{x_1,\dots, x_n\}$ taking values in a domain $D$ and a set of constraints $\calC$. Each constraint $C\in \calC$ is a function from $D^k$ to $\bbR^+$ applied to $k$ variables from $V$. Given an instance of a CSP, our goal is to assign values to the variables to maximize the total payoff of all constraints: $$\max_{x_1,\dots,x_n\in D^n} \sum_{C\in \calC} C(x_1,\dots, x_n).$$ Note, that we write $C(x_1,\dots, x_n)$ just to simplify the notation. In fact, $C$ may depend on at most $k$ variables. The parameter $k$ is called the arity of the CSP. In specific CSP problems, constraints $C$ come from a specific family of constraints. For example, in Max Cut, the domain is $D=\{-1,1\}$, and all constraints have the form $C(x_1,\dots,x_n) = \indicator(x_i\neq x_j)$; in Max 3LIN-2, the domain $D=\{0, 1\}$, and all constraints have the form $C(x_1,\dots,x_n) = \indicator(x_i\oplus x_j \oplus x_l = 0)$ or $C(x_1,\dots,x_n) = \indicator(x_i\oplus x_j \oplus x_l = 1)$. Various approximation algorithms have been designed for CSPs. The most basic among them, the ``trivial'' probabilistic algorithm simply assigns random values to the variables. It turns out, however, that in some cases this algorithm is essentially optimal. \citeasnoun{Hastad97} showed that for some CSPs e.g., 3LIN-2 and E3-SAT, beating the approximation ratio of the random assignment algorithm (by any positive constant $\varepsilon$) is NP-hard. Such problems are called approximation resistant. That is, a constraint satisfaction problem is approximation resistant, if for every positive $\varepsilon > 0$, it is NP-hard to find a $(A_{trivial}+\varepsilon)$ approximation, where $A_{trivial}$ is the approximation ratio of the random assignment algorithm. If there exists an algorithm with the approximation ratio $(A_{trivial}+\varepsilon)$ for some positive $\varepsilon$, we say that the problem {\em admits a non-trivial approximation}. It is still not known which constraint satisfaction problems are approximation resistant and which admit a non-trivial approximation. This is an active research question in approximation algorithms. Suppose now that in our instance of $k$-CSP, each variable is used by at most $B$ constraints. (For example, for Max Cut, this means that the maximum degree of the graph is bounded by B.) \citeasnoun{Hastad00} proved that such instances (which we call $B$-{\em bounded occurrence }$k$-CSPs) admit a non-trivial approximation. Let $\Opt$ denote the value of the optimal solution; $\Avg$ denote the expected value of the random assignment; and $\Alg$ denote the expected value returned by the algorithm. \citeasnoun{Hastad00} showed that there exists an approximation algorithm such that\footnote{The quantity $(\text{``value of the solution''} - \Avg)$ is called the {\em{advantage over random}}. The algorithm of \citeasnoun{Hastad00} is $O_k(B)$ approximation algorithm for the advantage over random: $$\Alg - \Avg \geq \frac{\Opt - \Avg}{O_k(B)}.$$} $$\Alg \geq \Avg + \frac{\Opt - \Avg}{O_k(B)}.$$ Here the hidden constant in $O_k(\cdot)$ may depend on $k$. \citeasnoun{Tre01} showed a hardness of approximation lower bound of $\Avg + (\Opt - \Avg)/(\Omega_k(\sqrt{B}))$. In this work, we study {\em{ordering}} constraints satisfaction problems. A classical example of an ordering $k$-CSP is the Maximum Acyclic Subgraph problem. Given a directed graph $G=(V,E)$, the goal is to find an ordering of the vertices $\pi: V\to \{1,\dots,n\}$ ($\pi$ is a bijection; $n=|V|$), so as to maximize the number of forward edges. In this case, the edges of the graph are constraints on the ordering $\pi$. An edge $(u,v)$ corresponds to the constraint $\pi(u)<\pi(v)$. Another example is the Betweenness problem. We are given a set of vertices $V$ and a set of constraints $\{C_{u,v,w}\}$. Each $C_{u,v,w}$ is defined as follows: $C_{u,v,w}(\pi) = 1$, if $u<v<w$ or $w<v<u$, and $C_{u,v,w}(\pi) = 0$, otherwise. The goal again is to find an ordering satisfying the maximum number of constraints. More generally, in an ordering $k$-CSP, each constraint $C$ is a function of the ordering that depends only on the relative order of $k$ vertices. The goal is given a set of vertices $V$ and a set of constraints $\calC$, to find an ordering $\pi: V\to [n]$ to maximize the total value of all constraints: $$\max_{\pi:V\to[n]} \sum_{C\in \calC} C(\pi).$$ Here $\pi$ is a bijection and $n=|V|$. If all constraints take values $\{0,1\}$, then the objective is simply to maximize the number of satisfied constraints. Note, that an {\em ordering} $k$-CSP is not a $k$-CSP. Surprisingly, we know more about ordering CSPs than about regular CSPs. \citeasnoun{GHM11} showed that every ordering CSP problem is approximation resistant assuming the Unique Games Conjecture (special cases of this result were obtained by \citeasnoun{GMR08} and \citeasnoun{CGM09}). On the positive side, \citeasnoun{BS90} showed that bounded degree Maximum Acyclic Subgraph, and thus every bounded occurrence ordering 2CSP, admits a non-trivial approximation. Their result implies that $\Alg \geq \Avg + (\Opt - \Avg)/O(\sqrt{B})$. \citeasnoun{CMM07} showed that a slight advantage over the random assignment algorithm can be also achieved for instances of Maximum Acyclic Subgraph ($\Alg \geq \Avg + (\Opt - \Avg)/O(\log n)$) whose maximum degree is not bounded. \citeasnoun{Gutin} showed that the ``advantage over the random assignment'' for ordering 3CSPs is {\em fixed--parameter tractable} (we refer the reader to the paper for definitions and more details). Finally, \citeasnoun{GZ12} proved that all bounded occurrence ordering 3CSPs admit a non-trivial approximation ($\Alg \geq \Avg + (\Opt - \Avg)/O_k(B)$). They also proved that there exists an approximation algorithm for {\em monotone }$k$-CSP (i.e., ordering CSPs, where all constraints are of the form $\pi(u_{i_1}) < \pi(u_{i_2}) < \dots < \pi(u_{i_k})$) with approximation ratio $1/k! + 1/O_k(B)$. {\textbf{Our results.}} We show that a very simple randomized local search algorithm finds a solution of expected value: \begin{equation}\label{eq:mainclaim} \Alg \geq \Avg + \frac{\Opt-\Avg}{O_k(B^{k+2})}. \end{equation} This algorithm works for every bounded occurrence ordering $k$-CSP. Consequently, all bounded occurrence ordering $k$-CSPs admit a non-trivial approximation. The running time of the algorithm is $O(n\log n)$. We do not know whether the dependence on $B$ is optimal. However, the result of \citeasnoun{Tre01} implies a hardness of approximation upper bound of $\Alg + (\Opt-\Avg)/\Omega_k(\sqrt{B}) \footnote{Every $k$-CSP can be encoded by an ordering $2k$-CSP by replacing every boolean variable $x$ with two variables $u_x^{\leftarrow}$ and $u_x^{\rightarrow}$, and letting $x = 1$ if and only if $\pi(u_x^{\leftarrow}) < \pi(u_x^{\rightarrow})$.}. \textbf{Techniques.} Our algorithm works as follows: first, it permutes all vertices in a random order. Then, $n$ times, it picks a random vertex and moves it to the optimal position without changing the positions of other vertices. We give an elementary proof that this algorithm performs better than the random assignment. However, the bound we get is exponentially small in $B$. Then, we improve this bound. Roughly speaking, instead of the original problem we consider the ``$D$-ordering'' problem, where the algorithm puts vertices in $D\approx Bk$ buckets (possibly, in a clever way), then it randomly permutes vertices in each of the buckets, and finally outputs vertices in the first bucket, second bucket, etc. This idea was previously used by \citeasnoun{CMM07}, \citeasnoun{GHM11}, \citeasnoun{Gutin} and \citeasnoun{GZ12}. The transition to ``$D$-orderings'' allows us to represent the payoff function as a Fourier series with relatively few terms. We prove that the $L_1$ weight of all coefficients of the payoff function is at least $\Avg + \Omega_k(\Opt-\Avg)$ (Note, that the optimal value of the ``$D$-ordering'' problem may be less than $\Opt$). Then we show that (a) for each vertex we can find one ``heavy'' Fourier coefficient $\hat{f}_{S}$; and (b) when the original local search algorithm moves a vertex it increases the value of the solution in expectation by at least $\Omega_k(\hat{f}_{S}/B)$. This concludes the proof. \textbf{Correction.} In the preliminary version of the paper that appeared at arXiv, we proved the main result of the paper, Theorem~\ref{thm:main}. We also gave an alternative, more complicated algorithm in the Appendix. We erroneously claimed that the performance guarantee of the alternative algorithm is slightly better than~(\ref{eq:mainclaim}). This is not the case. So the best bound known to the author is~(\ref{eq:mainclaim}). \pagebreak \section{Preliminaries} An instance of an ordering $k$-CSP problem $(V,\calC)$ consists of a set of vertices $V$ of size $n$, and a set of constraints $\calC$. An ordering of vertices $\pi: V \to \{1,\dots, n\}$ is a bijection from $V$ to $\{1,\dots, n\}$. Each constraint $C\in \calC$ is a function from the set of all ordering $\Sym_V = \{\pi: V \to \{1,\dots, n\}\}$ to $\bbR^+$ that depends on the relative order of at most $k$ vertices. That is, for every $C$ there exists a set $T_C\subset V$ of size at most $k$ such that if for two orderings $\pi_1$ and $\pi_2$, $\pi_1(u)< \pi_1(v) \Leftrightarrow \pi_2(u)<\pi_2(v)$ for all $u,v\in T_C$, then $C(\pi_1) = C(\pi_2)$. The value of an ordering $\pi$ equals $$\val(\pi, \calC) = \sum_{C\in \calC} C(\pi).$$ We will sometimes write $\val((u_1,\dots u_n), \calC)$ to denote the $\val(\pi, \calC)$ for $\pi: u_i\mapsto i$. We denote the optimal value of the problem by $\Opt(V,\calC)\equiv \max_{\pi\in \Sym_V} \val(\pi, \calC)$, the average value --- the value returned by the random assignment algorithm --- by $\Avg(V, \calC) = 1/n! \; \sum_{\pi\in \Sym_V} \val(\pi, \calC)$. \section{Algorithm} We now present the algorithm. \OpenFrame \textbf{Randomized Local Search Algorithm} \medskip \textbf{Input: }a set of vertices $V$, and a set of constraints $\calC$. \textbf{Output: }an ordering of vertices $(v_1,\dots,v_n)$. \begin{enumerate} \item Randomly permute all vertices. \item Repeat $n$ times: \begin{itemize} \item Pick a random vertex $u$ in $V$. \item Remove $u$ from the ordering and insert it at a new location to maximize the payoff. I.e., if $v_1, \dots, v_{n-1}$ is the current ordering of all vertices but the vertex $u$, then find a location $i$ that maximizes the $\val(v_1, \dots, v_{i-1}, u, v_{i+1},\dots v_{n-1}, \calC)$, and put $u$ in the $i$-th position. \end{itemize} \item Return the obtained ordering. \end{enumerate} \CloseFrame \begin{theorem}\label{thm:main} Given an instance $(V,\calC)$ of a $B$-bounded occurrence ordering $k$-CSP problem, the Randomized Local Search Algorithm returns a solution $\pi_{\Alg}$ of expected value \begin{equation}\label{eq:to-prove} \Exp\,\val (\pi_{\Alg}, \calC) \geq \Avg(V,\calC) + \frac{\Opt(V,\calC) - \Avg (V,\calC)}{O_k(B^{k+2})}. \end{equation} \end{theorem} \begin{remark} In fact, our proof implies a slightly stronger bound: The second term on the right hand side of the inequality~(\ref{eq:to-prove}) can be replaced with $(\Opt(V,\calC) - \Wst (V,\calC))/O_k(B^{k+2})$, where $\Wst(V,\calC)$ is the value of the worst possible solution. \end{remark} \begin{proof} I. We first show using an elementary argument that $$ \Exp\,\val (\pi_{\Alg}, \calC) \geq \Avg(V,\calC) + \alpha(B,k)(\Opt(V,\calC) - \Avg (V,\calC)), $$ for some function $\alpha(B,k)$ depending only on $B$ and $k$. This immediately implies that every bounded occurrence ordering $k$-CSP admits a non-trivial approximation. Then, using a slightly more involved argument we prove the bound~(\ref{eq:to-prove}). Observe, that the expected value of the solution after step 1 is exactly equal to $\Avg (V,\calC)$. So we need to estimate how much local moves at step 2 improve the solution. Let $\Delta_u$ be the maximum possible increase in the value of an ordering $\pi$, when we move $u$ to another position. In other words, $\Delta_u = \max_{\pi^+,\pi^-} (\val (\pi^+, \calC) - \val (\pi^-, \calC))$, where the orderings $\pi^+$ and $\pi^-$ differ only in the position of the vertex $u$. Let $\pi^*$ be the optimal ordering, and $\pi_*$ be the worst possible ordering. We can transition from $\pi^*$ to $\pi_*$ by moving every vertex $u$ at most once. Thus, $$\sum_{u\in V} \Delta_u \geq \val (\pi^*, \calC) - \val (\pi_*, \calC) = \Opt(V,\calC) - \Wst(V,\calC)\geq \Opt(V,\calC) - \Avg(V,\calC).$$ Now, our goal is to show that when the algorithm moves a vertex $u$, the value of the solution increases in expectation by at least $\alpha(B,k) \Delta_u$ for some function $\alpha$ depending only on $B$ and $k$. Fix a vertex $u$. Let $\pi^+$ and $\pi^-$ be the orderings that differ only in the position of the vertex $u$ such that $\Delta_u = \val (\pi^+,\calC) - \val (\pi^-, \calC)$. It may happen that the random permutation chosen by the algorithm at step 1 is $\pi^-$, and $u$ is chosen first among all vertices in $V$ at step 2. In this case, the algorithm can obtain the permutation $\pi^+$ by moving $u$, and thus it can increase the value of the solution by $\Delta_u$. However, the probability of such event is negligible. It is $1/n\cdot 1/n!$. The main observation is that the increase in the value of the ordering, when we move $u$, depends only on the order of the {\em neighbors }of $u$ i.e., those vertices that share at least one common constraint $C\in \calC$ with $u$ (including $u$ itself). We denote the set of neighbors by $N(u)$. Since each vertex participates in at most $B$ constraints, and every constraint depends on at most $k$ variables, $|N(u)|\leq kB$. Consider an execution of the algorithm. We say that $u$ is {\em fresh }if $u$ was chosen at least once in the ``repeat'' loop of the algorithm, and none of the neighbors were chosen before $u$ was chosen the first time. The probability that a variable $u$ is fresh is at least $\nicefrac{1}{2}|N(u)|^{-1}$. Indeed, the probability that at least one vertex in $N(u)$ is chosen is $1 - (1 - |N(u)|/n)^n > 1 - \nicefrac{1}{e}$; the probability that the first vertex chosen in $N(u)$ is $u$ is $1/|N(u)|$ (since all vertices in $N(u)$ have the same probability of being chosen first). If $u$ is fresh, then when it is chosen, its neighbors are located in a random order (since none of them was moved by the algorithm). Thus, with probability $1/N(u)!\geq 1/(kB)!$, the order of neighbors of $u$ is the same as in $\pi^-$. Then, by moving $u$ we can increase the value of the ordering by $\Delta_u$. Therefore, when the algorithm moves the vertex $u$, the value of the ordering increases in expectation by at least $$\Pr(u \text{ is fresh})\cdot \Pr (N(u) \text{ is ordered as }\pi^- \text{ after step 1})\cdot \Delta_u = \frac{\Delta_u}{2|N(u)|\,|N(u)|!}\geq \frac{\Delta_u}{kB\,(kB)!}.$$ This finishes the elementary proof that a positive $\alpha (B,k)$ exists. \medskip II. We now improve the lower bound on $\alpha (B,k)$. We show that for a fresh variable $u$, the value of the ordering increases in expectation by at least $\Omega_k(B^{-(k+1)})\Delta_u$, and thus (\ref{eq:to-prove}) holds. Let $L=\roundup{\log_2 (|N(u)|+1)}$ and $D=2^L$. Consider $D$ buckets $[D] =\{1,\dots, D\}$. For every mapping $x$ of the vertices to the buckets $v\mapsto x_v \in [D]$, we define a distribution $\calU_x$ on orderings of $V$. A random ordering from $\calU_x$ is generated as follows: put each vertex $v$ in the bucket $x_v$; then randomly and uniformly permute vertices in each bucket; and finally output vertices in the first bucket, second bucket, etc (according to their order in those buckets). Let $\calC_u$ be the set of constraints that depend on the vertex $u$. Since every variable participates in at most $B$ constraints, $|\calC_u|\leq B$. Let $f(x)$ be the expected total value of constraints in $\calC_u$ on a random ordering $\pi$ sampled from the distribution~$\calU_x$: $$f(x) = \Exp_{\pi\sim \calU_x}\Big[\sum_{C\in\calC_u} C(\pi)\Big].$$ Since the number of buckets $D$ is greater than or equals to $|N(u)|+1$, we may put every vertex in $N(u)$ in its own bucket and keep one bucket empty. Let $\pi^+$ and $\pi^-$ be the orderings as in part I of the proof: $\pi^+$ and $\pi^-$ differ only in the position of the vertex $u$, and $\val (\pi^+,\calC) - \val (\pi^-, \calC) = \Delta_u$. Consider mappings $x^+: V \to [D]$ and $x^-: V \to [D]$ that put only one vertex from $N(u)$ in every bucket and such that $x_v^+=x^-_v$ for every $v\neq u$, and $x^+$ orders vertices in $N(u)$ according to $\pi^+$, $x^-$ orders vertices in $N(u)$ according to $\pi^-$. For example, if $\pi^+$ arranges vertices in the order $(a,b, u, c)$, and $\pi^-$ arranges vertices in the order $(a,u, b, c)$, then $x^+ = (a\mapsto 1, *, b\mapsto 3, u\mapsto 4, c\mapsto 5)$ and $x^- = (a\mapsto 1, u\mapsto 2, b\mapsto 3, * , c\mapsto 5)$. Since the order of all vertices in $N(u)$ is fixed by $x^+$ and $x^-$, we have $f(x^+) = \val (\pi^+,\calC_u)$ and $f(x^-) = \val (\pi^-, \calC_u)$. Then \begin{eqnarray*} f(x^+) - f(x^-) &=& \val (\pi^+,\calC_u) - \val (\pi^-, \calC_u) \\ &=& \val (\pi^+,\calC) - \val (\pi^-, \calC) = \Delta_u. \end{eqnarray*} We now use Theorem~\ref{thm:bound-var}, which we prove in Section~\ref{sec:bound-var}. Let $X_v$ (for $v\in V$) be independent random variables uniformly distributed in $[D]$. By Theorem~\ref{thm:bound-var}, \begin{eqnarray*} \Exp[\max_{x_u \in D} f(x_u, \{X_v\}_{v\neq u}) - f(X_u, \{X_v\}_{v\neq u})] &\geq& \Omega_k(B^{-1}D^{-k})(f(x^+) - f(x^-))\\ &=& \Omega_k(B^{-(k+1)})\Delta_u. \end{eqnarray*} Here, $(x_u, \{X_v\}_{v\neq u})$ denotes the mapping $u\mapsto x_u$ and $v\mapsto X_v$ for $v\neq u$; and $(X_u, \{X_v\}_{v\neq u})$ denotes the mapping $v\mapsto X_v$ for all $v$. Observe, that when we sample random variables $X_v$, and then sample $\pi$ according to $\calU_{X}$, we get a random uniform ordering $\pi$ of all vertices in $V$. Thus, $$\Exp[f(X_u,\{X_v\}_{v\neq u})] = \Exp_{\pi \in \Sym_V}\Big[\sum_{C\in\calC_u} C(\pi)\Big] = \Exp_{\pi}[\val(\pi, \calC_u)].$$ Similarly, when we sample random variables $X_v$, set $x_u = \argmax_{x_u\in D}f(x_u, \{X_v\}_{v\neq u})$, and then sample $\pi'$ according to $\calU_{(x_u, \{X_v\}_{v\neq u})}$, we get a random uniform ordering of all vertices except for the vertex $u$. Denote by $LS(\pi, u)$ the ordering obtained from the ordering $\pi$ by moving the vertex $u$ to the optimal position. It is easy to see that if $\pi$ is a random uniform ordering, then $LS(\pi, u)$ has the same distribution as $LS(\pi', u)$, since the new optimal position of $u$ depends only on the relative order of other vertices $v$, and not on the old position of $u$. Hence, \begin{eqnarray*} \Exp[\max_{x_u \in D} f(x_u,\{X_v\}_{v\neq u})] &\equiv& \Exp_{\pi'}[\val(\pi', \calC_u)] \\ &\leq& \Exp_{\pi'}[\val(LS(\pi',u), \calC_u)]\\ &=& \Exp_{\pi}[\val(LS(\pi,u), \calC_u)]. \end{eqnarray*} Hence, \begin{eqnarray*} \Exp_{\pi}[\val(LS(\pi,u), \calC) - \val(\pi,u,\calC)] &=& \Exp_{\pi}[\val(LS(\pi,u), \calC_u) - \val(\pi,u, \calC_u)]\\ &\geq& \Omega_k(B^{-(k+1)})\Delta_u. \end{eqnarray*} \vskip -0.3mm \end{proof} \section{Theorem~\ref{thm:bound-var}}\label{sec:bound-var} \begin{theorem}\label{thm:bound-var} Let $D$ be a set of size $2^L$ (for some $L$). Consider a function $f: D^{n+1}\to \bbR$ that can be represented as a sum of $T$ functions $f_t: D^{n+1}\to \bbR$: $$f(x_0,x_1,\dots, x_n) = \sum_{t=1}^T f_t(x_0, x_1,\dots, x_n)$$ such that each function $f_t$ depends on at most $k$ variables $x_u$. Here, $x_0,\dots, x_n\in D$. Then, the following inequality holds for random variables $X_0,\dots, X_n$ uniformly and independently distributed in $D$: \begin{multline*} \Exp[\max_{x \in D} f(x,X_1,\dots, X_n) - f(X_0,X_1,\dots, X_n)] \geq \\ \Omega_k(T^{-1}|D|^{-k})\max_{x^{+}, x^{-},x_1,\dots, x_n \in D} (f(x^+,x_1,\dots, x_n) - f(x^-,x_1,\dots, x_n)). \end{multline*} \end{theorem} \begin{remark} The variable $x_0$ corresponds to $x_u$ from the proof of Theorem~\ref{thm:main}. The functions $f_t(x)$ are equal to $\Exp_{\pi\sim \calU_x} C_t(\pi)$, where $C_t$ is the $t$-th constraint from $\calC_u$. \end{remark} \begin{proof} Without loss of generality we assume that elements of $D$ are vertices of the boolean cube $\{-1,1\}^L$. We denote the $i$-th coordinate of $x\in D$ by $x(i)$. We now treat $f$ as a function of $(n+1)L$ boolean variables $x_u(i)$. We write the Fourier series of the function $f$. The Fourier basis consists of functions $$\chi_S(x_0,\dots, x_n) = \prod_{(u,i)\in S} x_u(i),$$ which are called {\em characters}. Each index $S\subset \{0,1,\dots,n\}\times \{1,\dots,L\}$ corresponds to the set of boolean variables $\{x_u(i) :(u,i)\in S\}$. Note, that $\chi_{\varnothing}(x_0,\dots, x_n) = 1$. The Fourier coefficients of $f$ equal $$\hat{f}_S = \Exp [f(X_0,\dots, X_n)\, \chi_S(X_0,\dots, X_n)],$$ and the function $f$ equals $$f(x_0,\dots, x_n) = \sum_{S} \hat{f}_S\, \chi_S(x_0,\dots, x_n).$$ \begin{remark} In the proof, we only use the very basic facts about the Fourier transform. The main property we need is that the characters form an orthonormal basis, that is, $$\Exp [\chi_{S_1}(X_0,\dots, X_n)\chi_{S_2}(X_0,\dots, X_n)] = \begin{cases} 1,&\text{if } S_1 = S_2;\\ 0,&\text{if } S_1\neq S_2.\\ \end{cases}$$ Particularly, for $S\neq\varnothing$, $$\Exp [\chi_{S}(X_0,\dots, X_n)] = \Exp [\chi_{S}(X_0,\dots, X_n)\chi_{\varnothing}(X_0,\dots,X_n)] = 0.$$ We will also need the following property: if $f$ does not depend on the variable $x_u(i)$, then all Fourier coefficients $\hat{f}_S$ with $(u,i)\in S$ are equal to 0. \end{remark} Here is a brief overview of the proof: We will show that the $L_1$ weight of Fourier coefficients of $f$ is at least $f(x^+,x_1,\dots, x_n) - f(x^-,x_1,\dots, x_n)$, and the weight of one of the coefficients $\hat{f}_{S^*}$ is at least $\Omega(T^{-1}|D|^{-k}(f(x^+,x_1,\dots, x_n) - f(x^-,x_1,\dots, x_n)))$. Consequently, if we flip a single bit $X_0(i^*)$ in $X_0$ to make the term $\hat{f}_{S^*} \chi_{S^*} (X'_0, X_1, \dots, X_n)$ positive, we will increase the expected value of $f$ by $|\hat{f}_{S^*}|$. Observe, that since each function $f_t$ depends on at most $kL$ boolean variables, it has at most $2^{kL} = |D|^k$ nonzero Fourier coefficients. Thus, $f$ has at most $T\,|D|^k$ nonzero Fourier coefficients~$\hat{f}_S$. Pick $x^{+}, x^{-},x^*_1,\dots, x^*_n$ that maximize $f(x^+,x^*_1,\dots, x^*_n) - f(x^-,x^*_1,\dots, x^*_n)$. We have $$ f(x^+,x^*_1,\dots, x^*_n) - f(x^-,x^*_1,\dots, x^*_n) = \sum_{S} \hat{f}_S (\chi_S(x^+,x^*_1,\dots, x^*_n) - \chi_S(x^-,x^*_1,\dots, x^*_n)). $$ If $S$ does not contain pairs $(0,i)$ corresponding to the bits of the variable $x_0$, then the character $\chi_S(x_0,x_1,\dots,x_n)$ does not depend on $x_0$, and $\chi_S (x^+,x^*_1,\dots, x^*_n) - \chi_S(x^-,x^*_1,\dots, x^*_n) = 0$, hence \begin{multline*} f(x^+,x^*_1,\dots, x^*_n) - f(x^-,x^*_1,\dots, x^*_n) =\\= \sum_{S: \exists i \text{ s.t. } (0,i)\in S} \hat{f}_S \cdot (\chi_S(x^+,x^*_1,\dots, x^*_n) - \chi_S(x^-,x^*_1,\dots, x^*_n)) \leq 2 \sum_{S: \exists i \text{ s.t. } (0,i)\in S} |\hat{f}_S|. \end{multline*} Pick a character $\hat{f}_{S^*}$ with maximum absolute value and pick one of the elements $(0,i^*)\in S^*$. Since the number of nonzero characters $\hat{f}_S$ is at most $T\,|D|^k$, $$|\hat{f}_{S^*}|\geq \frac{f(x^+,x^*_1,\dots, x^*_n) - f(x^-,x^*_1,\dots, x^*_n)}{2T\,|D|^k}.$$ Let $\sigma = \sgn(\hat{f}_{S^*})$. Define a new random variable $X_0'$, $$X_0'(i)= \begin{cases} X_0(i),&\text{for } i\neq i^*;\\ \sigma\, \chi_{S^*}(X_0,\dots,X_n)\, X_0(i),&\text{for } i = i^*. \end{cases} $$ Consider a character $\chi_S$. If $(0,i^*)\notin S$, then $\chi_S$ does not depend on the bit $x_0(i^*)$, hence $\Exp[\chi_S(X_0',X_1,\dots, X_n)] = \Exp[\chi_S(X_0,X_1,\dots, X_n)]$. On the other hand, if $(0,i^*)\in S$, then \begin{multline*} \Exp[\chi_S(X_0',X_1,\dots, X_n)]= \Exp\big[\frac{X'_0(i^*)}{X_0(i^*)}\,\chi_S(X_0,X_1,\dots, X_n)\big] =\\ \Exp[\sigma \chi_{S^*}(X_0,X_1,\dots, X_n) \chi_S(X_0,X_1,\dots, X_n)]= \begin{cases} 0,&\text{if } S\neq S^*;\\ \sigma,&\text{if } S = S^*. \end{cases} \end{multline*} The last equality holds because characters $\chi_S$ form an orthonormal basis. Therefore, $$\Exp[f(X'_0, X_1, \dots, X_n) - f(X_0, X_1, \dots, X_n)] = \sigma \hat{f}_{S^*} = |\hat{f}_{S^*}|.$$ We get \begin{multline*} \hskip -0.8mm\Exp[\max_{x \in D} f(x,X_1,\dots, X_n) - f(X_0,X_1,\dots, X_n)] \geq \Exp[f(X'_0, X_1, \dots, X_n) - f(X_0, X_1, \dots, X_n)] \\ = |\hat{f}_{S^*}| \geq \Omega_k(T^{-1}|D|^{-k})\max_{x^{*}, x_{*},x_1,\dots, x_n \in D} (f(x^+,x_1,\dots, x_n) - f(x^-,x_1,\dots, x_n)). \end{multline*} \end{proof} \pagebreak \section{Concluding remarks} We can guarantee that the algorithm finds a solution of value (\ref{eq:to-prove}) with high probability by repeating the algorithm $\Theta_k(B^{k+2})$ times (since the maxim possible value of the solution is $\Opt$). We note that our local search algorithm works not only for ordering $k$-CSPs, but also for (regular) $k$-CSPs. The algorithm first assigns random values to all variable $x_i$, and then, $n$ times, picks a random $i \in \{1,\dots, n\}$, and changes the value of the variable $x_i$ to the optimal value for fixed other variables. The approximation guarantee of the algorithm is $\Avg(V,\calC) + (\Opt(V,\calC) - \Avg (V,\calC))/O_{k,D}(B)$, here $k$ is the arity, and $D$ is the domain size of the CSP. The approximation guarantee has the same dependence on $B$ as the approximation guarantee of H{\aa}stad's (2000) original algorithm. The analysis relies on Theorem~\ref{thm:bound-var}.
\section{Introduction} With the recent detections of possible terrestrial planets in the habitable zones of other stars by the Kepler mission (Lissauer et al., 2012) a question arises: to what extent could these remote worlds support life similar to what we know here on Earth (terrestrial-type life)? There are many conditions for a planet to fulfil to support terrestrial-type life, which can be controlled by either geological, climatological, orbital, or geophysical processes. In this study, the first in a series, we focus on the dynamical factors that may affect the habitability of a terrestrial exoplanet. We identify several key features that we believe are important, if not essential, for supporting terrestrial-type life. These are i) a stable climate, ii) low to moderate ($<$ 50$^\circ$) seasonal temperature variations, iii) low to moderate diurnal temperature variations, and iv) low to moderate spatial variation of temperature over the planet. These conditions need to be supplemented with the following: v) regular, low-amplitude obliquity oscillations, vi) moderate obliquity to induce seasonal variations, vii) moderate rotation rate, and viii) small orbital perturbations. We examine each of these below.\\ For the current Earth, the first condition requires a constant obliquity and a low eccentricity. Earth's life is both land-based and water-based and requires a stable climate over millions or even billions of years. In addition to continental drift the long-term climate on the Earth is largely regulated by the Milankovi\'{c} cycles (Milankovi\'{c}, 1941). These cycles are related to variations in Earth's orbit and their influence on the obliquity. The main cycles are related to its equatorial precession (with period 26~kyr, corresponding to a frequency $\dot{\psi}=$-50.5~''/yr, where $\psi$ is the angle between the vernal equinox and the intersection of the equator with the ecliptic), the obliquity variations of the Earth (with period 41~kyr, corresponding to the frequency $\dot{\psi}-s_3$, where $s_3$ is the Earth's nodal eigenfrequency), and the orbital eccentricity (with periods 100~kyr and 400~kyr, corresponding to frequencies $g_2+g_5$ and $g_2-g_5$, where $g_2$ and $g_5$ are the eccentricity eigenfrequencies of Venus and Jupiter respectively). The present obliquity variations of the Earth are small: the amplitude is just 1.5$^\circ$, and the low amplitude is a result of the presence of the Moon. However, even these small oscillations are enough to cause regular ice ages through the positive feedback effect (Deser et al., 2000).\\ The second condition that is necessary to make the planet desirable for terrestrial-type life leans on the following. Spiegel et al. (2009) have shown that the polar regions of a planet with perpendicular spin are mostly uninhabitable because of the lack of seasons: the cold regions will mostly stay cold, possibly even below freezing. On Earth the onset of the ice ages tends to favour low obliquity for two reasons: the reduction in overall summer insolation at high latitudes and the corresponding reduction in mean insolation (Huybers \& Wunsch, 2005). The cooling would result in more ice building up near the polar caps, increasing the albedo which then instigates a positive feedback effect. However, there are other theories that show that precession and eccentricity forcing on the insolation also play a role (e.g. Imbrie \& Imbrie, 1980; Paillard, 1997; Huybers, 2011) and the debate is ongoing. The weak seasonal effects at low obliquity and the corresponding constant low temperatures at high latitudes may imply that habitability is increased for planets that have a moderate obliquity because the associated seasons make the polar regions partly habitable by increasing the yearly mean insolation, and thus the temperature. However, too high an obliquity may not be favourable for habitability either (Williams \& Kasting 1997; Williams \& Pollard, 2003; Spiegel et al., 2009) because the long summers at the poles could cause large seasonal temperature variations above large continents (Williams \& Pollard, 2003) and the highest temperatures could exceed 50~$^\circ$~C. Additionally the long periods of darkness even at mid latitudes could cause difficulties for photosynthetic life to survive. Thus it appears that the current obliquity favours the development of terrestrial-type life rather than more extreme values.\\ The third condition, a low to moderate diurnal temperature variation, is regulated by a sufficiently slow rotation rate, a thick atmosphere and oceans. Diurnal temperature changes are partially governed by the relaxation time scale for atmospheric heat losses and temperature variations. On the Earth this time scale is 20 days (Matsuda, 2000) and the diurnal temperature variation is approximately 10~K. For Mars the story is very different: the relaxation time is comparable to its rotation period and the diurnal temperature variation is approximately 60~K (Matsuda, 2000). Part of the reason the diurnal variation on Mars is much higher than on Earth is because the latter's oceans store heat much more effectively than land and Mars' atmosphere is much more tenuous. However, a very slow rotation will cause the Hadley cell to cover the whole planet (Farrell, 1990) and the diurnal temperature variations may be diminished. A fast rotation, on the other hand, would decrease the heat transport as argued in the next paragraph. By themselves these temperature changes may not pose a problem but, coupled with the seasonal variations, they may be difficult for land-based life to adapt to. \\ The fourth condition, low to moderate spatial variation of temperature over the planet, can be the result of the following. Simulations of global circulation models on an Earth-like planet with different rotation rates have shown that for a fast rotating Earth the atmosphere experiences the creation of many small Hadley-like cells (Williams, 1988a) and most of the heat transport is caused by baroclinic eddies. These eddies decrease the efficiency of heat transportation from the warmer regions to the colder regions. This possible reduction in heat transportation by the atmosphere may decrease the habitability of the planet because the polar (and thus colder) regions would probably stay cold (Williams, 1988a). However, even though there appears evidence for a more equable climate having existed during the Cretaceous and Eocene aeons (e.g. Barron, 1989), which may have been caused by latitudinal ocean currents rather than the present-day longitudinal ones (Bice \& Marotzke, 2002), it is not clear to what extent the size of the Hadley cells depends on rotation rate or on land mass distribution, and how this increased equator to poleward heat flow could have been sustained (Barron, 1983). Nevertheless it is probable the area of habitability is smaller on a fast-rotating planet than it is on a slow-spinning one. Secondly, we have the issue that pertains to the effect of tides on ocean mixing. It is well known that tidal forcing causes the mixing of stratified layers in the ocean (e.g. Egbert \& Ray, 2000), though recent work shows that mixing by swimming sea creatures could be equally important (Katija \& Dabiri, 2009). Whatever its source this mixing has profound effects on the Earth's climate because it allows for more efficient energy exchange between the atmosphere and the ocean, when cooler water reaches the surface from the deeper regions. The mixing also aids in the transport of water from warmer regions to colder ones by currents such as the Gulfstream (Garrett, 2003). Last, the mixing brings up important nutrients from the depths of the ocean that micro-organisms residing closer to the surface can feed on. Thus, it appears that in a simplistic sense ocean mixing could be an important ingredient in the sustenance of terrestrial-type life, and the habitability of the Earth might be decreased without this effect.\\ Regarding the second set of conditions for the existence of terrestrial-type life, v), vi) and vii) are satisfied because of the presence of the Moon, while conditions v) and viii) are the result of the planets all having small eccentricities and mutual inclinations, and the Earth being far from a secular orbital resonance and from a (secular) spin-orbit resonance. Thus at first glance it appears that the Moon plays a key role in supporting life on Earth. However, some of the above definitions should be interpreted with care. Mars appears to fulfil several of these criteria (moderate rotation speed, moderate obliquity, moderate orbital perturbations, moderate seasonal temperature variations). In addition, it could have a stable spin-axis if the secular frequencies where different. However, the slow rotation comes from the fact that it is a planetary embryo (Dauphas \& Pourmand, 2011), and embryos appear to be less adapted for life because they lack the gravitational strength to keep a thick atmosphere and oceans against Jeans escape, impact erosion or stellar wind pick-up over 4.5 Gyrs. The mass of an embryo may be regulated by an `isolation mass', which is one order smaller than an Earth mass at 1~AU. Thus for the remainder of this paper we focus on a fully-formed planet with a satellite, such as the Earth-Moon system. But is the Earth-Moon system unique in its architecture or is the current system a common outcome? We aim to answer that question and relate the findings to the habitability of the resulting systems. \\ It is generally believed that the Moon formed through the giant impact of a Mars-sized body with the proto-Earth (e.g. Hartmann \& Davis, 1975; Cameron \& Ward, 1976; Ida et al, 1997; Kokubo et al., 2000; Canup, 2004). This impact melted the impactor and part of the proto-Earth, and resulted in much of the ejecta orbiting the Earth (Benz et al., 1991; Cameron, 1997; Canup, 2004). The Moon subsequently accreted from this disc of debris in less than a year (Ida et al., 1997). Usually the final disc mass was less than 2.5 lunar masses and orbited inside the Earth's Roche limit, currently at 2.9 Earth radii $(R_{\oplus})$. The Moon ends up orbiting in the Earth's equatorial plane. Tidal evolution over the next 4.5~Gyr pushed it towards its current orbit (Goldreich, 1966; Touma \& Wisdom, 1994). With the aid of numerical simulations Elser et al. (2011) conclude that binary planetary systems, such as the Earth-Moon system, occur approximately 8\% of the time, ranging from 2.5\% up to 25\%. However, their studies did not take subsequent tidal evolution into account.\\ Goldreich (1966) and Touma \& Wisdom (1994) have shown that the Earth's current obliquity of 23.5$^\circ$ requires its obliquity to have been approximately 10$^\circ$ just after impact. Similarly, Earth's rotation period was approximately 5~h after the impact and the Moon's inclination with respect to the Earth's equator was about 10$^\circ$. The initial spin of the Earth being nearly perpendicular to its orbital plane is a fairly rare outcome after a Moon-forming impact: it has been shown that the obliquities of the terrestrial planets are isotropically distributed (Chambers, 2001; Kokubo \& Ida, 2007) and their resulting rotation rates are approximately half of their maximum values (Kokubo \& Genda, 2010). It is likely that the planetary rotation period is partially determined by the impact that formed the satellite (Lissauer, 2000) and the satellite mass is related to the impact parameter of the collision: grazing impacts produce heavier satellites and speed up the planet's rotation rate much more than head-on collisions (Kokubo et al., 2000). This result suggests there is a relation between the planet's rotation period and the satellite that has formed. However, as we show in Appendix A, the relation is not one-on-one but rather a rough estimate. Additionally, an isotropic obliquity distribution favours coplanar spins rather than perpendicular ones. What is then the evolution of the Earth-Moon system if the Earth had been much more oblique just after the Moon-forming impact?\\ Atobe \& Ida (2007) studied the tidal evolution of systems consisting of an Earth-like terrestrial planet with a satellite ranging from sub-lunar to super-lunar mass on a circular orbit. They varied the initial mass of the satellite ($m_s$) and the initial obliquity of the planet ($\varepsilon_0$) but kept the initial semi-major axis of the satellite at 3.8~$R_{\oplus}$, the Earth's rotation period at 5~h and the satellite's inclination with respect to the Earth's orbit, $i$, at 1$^\circ$. They showed that there are essentially three outcomes when these systems are evolved towards completion: i) the satellite recedes from the planet and approaches (and sometimes reaches) the outer co-rotation radius (case A), ii) the satellite first recedes from and then approaches the planet resulting in collision (case B), or iii) the satellite first recedes from the planet and then approaches it but gets caught at the inner co-rotation radius (case C). In general terms, case A occurs for low-mass satellites at low prograde or high retrograde obliquities. Case B occurs for initial obliquities $\varepsilon_0 \in (\sim 60^\circ, \sim 120^\circ)$ and case C occurs for heavy satellites ($m_s\gtrsim 0.03\,m_p$, with $m_p$ being the planetary mass) at low obliquity. We have summarised the results in Fig.~\ref{atobeida}, taken from Atobe \& Ida (2007), where the symbol $\gamma_0$ is $\varepsilon_0$ in our notation. The outcome case A is further subdivided according to whether the system is still evolving. For case A1 prograde planetary spins will have the planetary obliquity, $\varepsilon_p$ evolve to $\varepsilon_p \rightarrow 0^\circ$ while retrograde spins will evolve to $\varepsilon_p \rightarrow 180^\circ$. In the outcome A2 the system has stopped evolving. Outcome A3 is a particular case of A1, where a retrograde-spinning planet will eventually spin prograde due to the dominance of solar tides over satellite tides (for A1 this does not happen). \\ From this figure it appears that the outcome for oblique planets can be very different from the current Earth-Moon system. Atobe \& Ida (2007) studied the evolution of these systems until their ultimate state. For light satellites, such as the Moon, the time scale to reach the final configuration is much longer than the age of the Solar System, exceeding the Hubble time for masses below half a lunar mass. Thus their results cannot be used to directly determine the state of the system at the current epoch i.e. 4.5~Gyr after its formation.\\ \begin{figure} \resizebox{\hsize}{!}{\includegraphics[]{atobeida.eps}} \caption{This figure shows the regions of the three possible outcomes of planetary motion as a function of satellite mass and original planetary obliquity. Taken from Atobe \& Ida (2007). See text for details.} \label{atobeida} \end{figure} In this paper we determine the state of fictitious systems consisting of an Earth-like terrestrial exoplanet and a satellite whose mass ranges from sub-lunar to super-lunar values. We investigate the tidal evolution of these systems starting with different initial planetary obliquities and satellite masses. We do this by performing a series of simulations similar to those done by Atobe \& Ida (2007) but we cease them after the system has reached an age of 4.5~Gyr. We then analyse the states of each system and determine which regions of the phase space are the most suitable for supporting terrestrial-type life. In addition we attempt to determine how likely it is to produce a system similar to our own from a range of initial conditions. This paper is structured as follows.\\ In the next section we lay out the equations of motion consisting of the star, planet and satellite. These are based on the work of Bou\'{e} \& Laskar (2006) and Correia (2009). In section 3 we describe our numerical methods and compare the evolution from our simulations to earlier results presented in Bou\'{e} \& Laskar (2006) for the conservative motion, and Touma \& Wisdom (1994) for the tidal evolution. In section 4 we present the results of our numerical simulations, focusing on the final states of the system and how these may change with slightly different initial conditions and passages through secular spin-orbit resonances. In the last section we present a summary and conclusions. \section{Theory of precession and tidal evolution} We study the evolution of a hierarchical system consisting of a central star with mass $M_*$, a terrestrial planet with mass $m_p$ and a satellite with mass $m_s$. The planet and satellite are oblate spheroids with zonal harmonics $J_2 = (C-A)/m_pR^2$, where $R$ is the equatorial radius of the body under consideration and $C$ and $A$ are the principal and secondary moments of inertia. The planet and satellite rotate about the axis with maximal inertia, $C$. From now on, we shall use the dimensionless variant $\mathcal{C}=C/m_pR^2$. We follow Bou\'{e} \& Laskar (2006) and Correia (2009) in the derivation below.\\ The conservative motion of the system can be tracked through the conservation of the total angular momentum \begin{equation} {\bf \dot{L}_p + \dot{L}_s +\dot{H}_p + \dot{H}_s} = 0, \label{angmomcons} \end{equation} where a dot denotes a time derivative, ${\bf H}_i = \mathcal{C}_im_iR_i^2\nu_i{\bf s}_i$ are the spin angular momenta of the planet and satellite and ${\bf L}_i=m_in_ia_i^2(1-e_i)^{1/2}{\bf k}_i$ are their orbital angular momenta. Here $\nu_i$ is the spin frequency of either body, $a_i$ are the semi-major axes, $e_i$ are their eccentricities and $n_i$ are their mean motion. The vectors ${\bf s}_i$ are the unit vectors in the direction of the spin angular momentum of the planet and the satellite while the vectors ${\bf k}_i$ are the unit vectors of the orbital angular momenta. In what follows the subscript $i$ can refer to either the planet or the satellite. Tidal evolution acts on time scales much longer than the orbital period of both the planet and the satellite, so equation (\ref{angmomcons}) is averaged over the mean anomalies of both bodies and the periapse of the planet-satellite pair. The resulting equations for the conservative motion are then given by (Bou\'{e} \& Laskar, 2006) \begin{eqnarray} \label{cons} {\bf \dot{H}}_i & = & -\alpha_i \cos \varepsilon_i \, {\bf k_p \times s_i} -\beta_i \cos \theta_i \, {\bf k_s \times s_i}, \nonumber \\ {\bf \dot{L}}_s & =& -\gamma \cos I \, {\bf k_p \times k_s} + \sum_i \beta_i \cos \theta_i \, {\bf k_s \times s_i}, \\ {\bf \dot{L}}_p & =& \gamma \cos I \, {\bf k_p \times k_s} + \sum_i \alpha_i \cos \varepsilon_i \, {\bf k_p \times s_i}. \nonumber \end{eqnarray} The equations above assume that the eccentricity of the satellite is constant during one full rotation of the line of apses. This does not necessarily have to be true, but we shall assume that it is so in this simplified model. A study that also accounts for the eccentricity evolution of the satellite is left for the future. \\ In equations (\ref{cons}) above we have introduced the angles $I = \arccos({\bf k_p \cdot k_s})$, which is the inclination of the orbit of the satellite with respect to the planet's orbital plane, $\varepsilon_i = \arccos({\bf k_p \cdot s_i})$ are the obliquities of both bodies with respect to the planet's orbital plane and $\theta_i = \arccos({\bf k_s \cdot s_i})$ are the obliquities of both bodies with respect to the satellite's orbital plane. In this study we are primarily interested in the quantities $I$, $\varepsilon_p$ and partially $\theta_s$. We also introduced the precession constants \begin{eqnarray} \alpha_i &=& \frac{3GM_*m_iJ_{2i}R_i^2}{2a_p^3(1-e_p^2)^{3/2}}, \nonumber \\ \beta_i &=& \frac{3Gm_pm_sJ_{2i}R_i^2}{2a_s^3(1-e_s^2)^{3/2}}, \\ \gamma &=& \frac{3GM_*m_sa_s^2(2+3e_s^2)}{8a_p^3(1-e_p^2)^{3/2}}. \nonumber \end{eqnarray} Roughly, the quantities $\alpha_i$ are the precession rates of the figures of the planet and satellite caused by the torques from the Sun, the quantities $\beta_i$ are the precession rates of the poles of both bodies caused by mutual torques between them, and $\gamma$ is the precession rate of the nodes of the satellite's orbit on the planet's orbit.\\ On long time scales the system is subject to tidal forces, most notably between the planet and the satellite. In this study we adopt the constant time delay model of Mignard (1979, 1980) for its simplicity, keeping the eccentricity of the planet and satellite constant (but not necessarily zero). The tidal forces from the planet and the satellite on each other act to reduce their rotation rates. The reduction in spin angular momentum is compensated by an increase in the orbital angular momentum of the satellite. The averaged equations of motion governing the tidal evolution of the planet-satellite pair are (Correia, 2009) \begin{eqnarray} {\bf \dot{H}}_i &=& -{\textstyle{\frac{1}{2}}}K_i[X^{-6,0}_0(e_s)({\bf s}_i + \cos \theta_i {\bf k}_s)\nu_i \nonumber \\ &-& 2n_sX^{-8,0}_0(e_s)(1-e_s^2)^{1/2}{\bf k}_s], \\ {\bf \dot{L}}_s &=& {\textstyle{\frac{1}{2}}}\sum_iK_i[X^{-6,0}_0(e_s)({\bf s}_i + \cos \theta_i {\bf k}_s)\nu_i \nonumber \\ &-& 2n_sX^{-8,0}_0(e_s)(1-e_s^2)^{1/2}{\bf k}_s], \nonumber \end{eqnarray} where we introduced \begin{eqnarray} K_p = 3k_{2p}Gm_s^2R_p^5\Delta t_p a_s^{-6}, \nonumber \\ K_s = 3k_{2s}Gm_p^2R_s^5\Delta t_s a_s^{-6}. \end{eqnarray} Here $k_{2i}$ is the secular Love number of the planet or satellite, $\Delta t_i$ is the tidal delay and $X^{-6,0}_0(e)$ and $X^{-8,0}_0(e)$ are Hansen coefficients. These are defined from \begin{equation} \Bigl(\frac{r}{a}\Bigr)^n\exp({\rm i}mv) = \sum_{l=-\infty}^{\infty}X_l^{n,m}(e)\exp({\rm i}lM), \end{equation} where $r$ is the distance of the planet to the star, $v$ is the true anomaly, $M$ is the mean anomaly, $l$, $m$ and $n$ are integers and ${\rm i}^2=-1$ is the imaginary unit. The time delay is related to the tidal dissipation parameter, $Q$, via $\Delta t = (Q\nu)^{-1}$.\\ The planet-satellite pair is not isolated and it will be subjected to tidal forces from the central star. For simplicity we only incorporated the tidal action from the star on the planet, and thus the reduction in the planet's spin angular momentum is compensated by an increase in its orbital angular momentum, giving \begin{eqnarray} {\bf \dot{H}}_p &=& -{\textstyle{\frac{1}{2}}}N[X^{-6,0}_0(e_p)({\bf s}_p + \cos \varepsilon_i {\bf k}_p)\nu_p \nonumber \\ &-& 2n_pX^{-8,0}_0(e_p)(1-e_p^2)^{1/2}{\bf k}_p], \end{eqnarray} with $N=3k_{2p}GM_*^2R_p^5\Delta t_p a_p^{-6}$. As the satellite recedes from the planet and the rotation of both bodies slows down, their $J_2$ coefficients decrease because the rotational deformation of their figures becomes less severe. The $J_2$ values of the planet and satellite are updated according to $J_{2i} = \frac{1}{3}k_{si}R_i^3\nu_i^2/Gm_i$ (Atobe \& Ida, 2007), where $k_s$ is the secular Love number, which is assumed to be constant. \\ Now that we have discussed the basic theory of the tidal and conservative motion, we describe our numerical methods below and compare them with previous results. \section{Numerical methods and tests} In order to determine the evolution of the planet-satellite system with various initial conditions, we integrated the equations of motion with the aid of computer codes. We wrote three versions. The first only integrates the conservative motion without any tidal evolution. A second version only includes planet-satellite tides while the third version also includes the solar tides. The integration was performed using the Bulirsch-Stoer method (Bulirsch \& Stoer, 1966). We checked our code against the results of Touma \& Wisdom (1994) and Bou\'{e} \& Laskar (2006), which were all reproduced. \\ The input parameters are the masses, obliquities, semi-major axes, eccentricities, radii, moments of inertia $\mathcal{C}$, tidal dissipation factors $Q$, Love numbers $k_2$, mutual inclination $i$, and rotation periods. The values of $J_2$ were derived from the input values. The unit vectors were computed from \begin{eqnarray} \bf{k}_p &=& (0,0,1)^T, \\ \bf{k}_s &=& (\sin I \sin \Omega, -\sin I \cos \Omega, \cos I)^T, \\ \bf{s}_s &=& (\sin \varepsilon_p \sin \psi, -\sin \varepsilon_p \cos \psi, \cos \varepsilon_p)^T, \end{eqnarray} with $\bf{s}_s$ generated similar to $\bf{s}_p$. For a close satellite we set $\Omega = \psi$ and $I \sim \varepsilon_p$ so that $i \sim 0$. In Table~\ref{tab1} we have listed all the input parameters and their initial values.\\ It is well known that integrating the tidal equations backward in time using the current tidal dissipation rate causes the Moon to fall onto the Earth approximately 1~Gyr ago (Touma \& Wisdom, 1994), so we scaled the final integration time to be 4.5~Gyr to obtain the current system state. In most cases we also lowered $Q_p$ to 10 to hasten the evolution and save CPU time. We made sure that this did not affect the final outcome. \\ \begin{table} \begin{tabular}{ccc} Quantity & Planet & Satellite \\ \hline \\ $m_i$ & $3.005 \times 10^{-6}$~$M_\odot$ & 0.0025 - 0.05 $m_p$ \\ $R_i$ & 4.2634965 $\times 10^{-5}$~AU & 1.161848 $\times 10^{-5}$~AU\\ $P_i$ & 5~h (3~h and 7~h) & 18.7~h \\ $\varepsilon_i$ & 0$^\circ$-175$^\circ$ & variable\\ $\theta_i$ & variable & 0 \\ $\mathcal{C}$ & 0.331 & 0.394\\ $k_2$ & 0.300 & 0.02264 \\ $k_s$ & 0.95 & 1.2 \\ $Q_i$ & 20 & 30 \\ $a_i$ & 1.0~AU (0.75~AU and 1.25~AU) & 3.8~$R_\oplus$\\ $e_i$ & 0 & 0\\ $i$ & - & 0, 1$^\circ$ and 5$^\circ$ \end{tabular} \caption{Initial values of various quantities for our numerical simulations.} \label{tab1} \end{table} To demonstrate the validity of the code and to show the different types of precession without tides, we plot the evolution of a fictitious Earth-Moon system for various lunar distances in Fig.~\ref{em}. We have taken the current system and simply decreased the lunar semi-major axis, keeping everything else the same. Thus the evolution at short lunar semi-major axis is unlikely to be representative of the past lunar orbit but serves to illustrate the various types of motion of the system. \\ The top two panels are for the current system with $a_s = 60$~$R_\oplus$, the middle panels have $a_s = 10$~$R_\oplus$ and the bottom panels are for $a_s = 5$~$R_\oplus$. The left panels plot the inclination of the lunar orbit with respect to the ecliptic ($I$) in red, the inclination of the lunar orbit with respect to the Earth's equator ($i$) in blue, and the obliquity of the Earth with respect to the ecliptic ($\varepsilon_p$) in green. In the right panels we plotted the evolution of the Moon's nodes on the ecliptic ($\Omega$) in red, and the node of the Earth's equator on the ecliptic ($\psi$) in green. We distinguish three types of motion, as outlined in Atobe \& Ida (2007). On the top panels, the Moon's orbit and the Earth's spin precess around $\bf{k}_p$, which results in $I$ and $\varepsilon$ being nearly constant, and $i$ oscillating between $\varepsilon_p - I$ and $\varepsilon_p+I$ with a period equal to the revolution of $\Omega$ ($\sim$ 18.5~yr). On the bottom panels, the Moon's orbit precesses about the common angular momentum vector of the spin of the Earth and the lunar orbit, $\bf{s}_p + \bf{k}_s$. Thus, the mutual inclination, $i$, stays almost constant while $I$ and $\varepsilon_p$ oscillate out of phase with a period equal to half the precession period of the satellite's orbit on the equator of the planet (e.g. Kinoshita \& Nakai, 1991). Superimposed on this is a long-period precession of $\bf{k}_s + \bf{s}_p$ (bottom-right panel). In the middle panels, the precession of both $\bf{s}_p$ and $\bf{k}_s$ is neither around their sum nor around $\bf{k}_p$, but somewhere in between (Atobe \& Ida, 2007). Thus $I$, $i$ and $\varepsilon_p$ all oscillate with large amplitude. Here we find that the precession period of the Earth's axis is about 500~yr; Bou\'{e} \& Laskar (2006) found it was 450~yr.\\ \begin{figure*} \resizebox{\hsize}{!}{\includegraphics[angle=-90]{em.eps}} \caption{Orbital evolution of the Earth-Moon system for several different initial configurations. The top panels depict the current state with the Moon at 60~$R_{\oplus}$. The middle panels pertain to the Moon being at 10~$R_{\oplus}$ and the bottom panels have the Moon at 5~$R_{\oplus}$. On the left column we plot $I$ (red), $\varepsilon_p$ (green) and $i$ (blue). The right panels plot $\Omega$ (red) and $\psi$ (green).} \label{em} \end{figure*} The tidal evolution of the Earth-Moon system to the present is shown in Fig.~\ref{emtides}, with initial $m_s = 0.01223$~$m_{\oplus}$, $\varepsilon_p = 7.3^\circ$ and $I=19^\circ$. The values of the other relevant quantities are in Table~\ref{tab1}. The top-left panel depicts the obliquity of the Earth (green) and the inclination of the Moon with respect to the ecliptic (red) versus the semi-major axis of the lunar orbit in Earth radii. To reproduce the current lunar ecliptic inclination, a high original inclination close to the Earth is needed, $i \sim 11^\circ$ (Touma \& Wisdom, 1994). This appears in contradiction with lunar formation theory from an impact (Canup, 2004). A possible solution to this problem is if the Moon passed through the evection and eviction resonances at 4.6~$R_\oplus$ and 6~$R_\oplus$ (Touma \& Wisdom, 1998), or maybe if the Earth originally had two moons (Jutzi \& Asphaug, 2011). \\ The top-right panel of Fig.~\ref{emtides} depicts the rotation period of the Earth vs the semi-major axis of the Moon. The results of both the top panels agree with those presented in Touma \& Wisdom (1994). The bottom-left panel depicts the semi-major axis of the lunar orbit with time. It is well known that the current dissipation in the Earth is anomalously high (e.g. Williams, 1999), caused by a near-resonance between ocean free modes and tidal forcing (Webb, 1982). Here we have just scaled the time such that the current system is reproduced at 4.5~Gyr. Unlike the tidal model with constant phase lag, where $Q$ is independent of the tidal forcing frequency, for the Mignard tidal model the semi-major axis evolution is not expressible in closed form as a function of time. At small semi-major axis $a_s(t) \propto t^{0.15}$ and for large semi-major axis $a_s(t) \propto t^{0.12}$, similar to, but not the same as for the constant phase lag model where $a_s(t) \propto t^{2/13} \approx t^{0.153}$. In the same figure, the bottom-right panel depicts $I$ and $\varepsilon_p$ as functions of time: $I$ decreases monotonically to about 5$^\circ$ while $\varepsilon_p$ continues to increase.\\ \begin{figure*} \resizebox{\hsize}{!}{\includegraphics[angle=-90]{emtides.eps}} \caption{Tidal evolution of the Earth-Moon system. This plot should be compared with Touma \& Wisdom (1994). The top-left panel plots $I$ (red) and $\varepsilon_p$ (green) vs $a_M$. The top-right panel shows the Earth's rotation period vs $a_M$. The bottom-left panel shows $a_M$ vs time and the bottom-right panel depicts $I$ (red) and $\varepsilon_p$ (green) vs time.} \label{emtides} \end{figure*} The result of a final test is depicted in Fig.~\ref{emprec}, which plots the precession frequency of the Earth's spin (black line) and the corresponding precession period (grey line) as a function of lunar distance. The data points were generated from numerical simulations: we tidally evolved the Earth-Moon system up to a pre-determined value of the semi-major axis of the Moon, using the initial conditions of Table~\ref{tab1} with initial $\varepsilon_p = 7.3^\circ$ and $I=19^\circ$, similar to Touma \& Wisdom (1994). Once the Moon had evolved to the designated distance, the simulation was stopped and the final state was integrated for 100~kyr without tides. The general shape and magnitude of the curve agrees with Touma \& Wisdom (1994) and Bou\'{e} \& Laskar (2006), but the maximum is at a larger lunar semi-major axis. We believe this is caused by us changing the value of the Earth's $J_2$ while Bou\'{e} \& Laskar (2006) kept it constant. \\ \begin{figure} \resizebox{\hsize}{!}{\includegraphics[angle=-90]{emprec.eps}} \caption{Precession frequency, $\dot{\psi}$ (black line) and the period of revolution of the Earth's spin axis (grey line) as a function of lunar separation.} \label{emprec} \end{figure} Now that we have demonstrated the validity of our codes through comparison with earlier work, we performed a series of simulations similar to Atobe \& Ida (2007): we fixed all the input quantities of the system to their values listed in Table~\ref{tab1}, but changed the initial obliquity of the planet and the mass of the satellite. The obliquity was increased from 0 to 175$^\circ$ in steps of 5$^\circ$ while the satellite mass ranged from 0.0025~$m_p$ to 0.05~$m_p$ in steps of 0.0025~$m_p$. Each of these planet-satellite systems were integrated for 4.5~Gyr and their endstates were recorded. These endstates were subsequently integrated for 100~kyr without tides to determine the spin precession frequency of the planet and compare them with analytical results of Bou\'{e} \& Laskar (2006). These endstates should quantitatively reflect the possible configurations of the planet-satellite system for different values of the initial satellite mass and initial planetary obliquity. While the outcome is similar to Atobe \& Ida (2007), our approach is different because we terminate the simulations at a given time rather than determine the final state of the system after all tidal evolution. We also checked whether the planet's precession frequency will pass through a secular spin-orbit resonance with one or more of the inclination eigenfrequencies of the solar system, $s_i$ (Brouwer \& van Woerkom, 1950). Such a resonance would induce a large, sudden increase in the planet's obliquity. We saved the state of the system whenever such a resonance was crossed for future analysis. \section{Results} In this section we present the results from our numerical simulations. We have used the third version of our code i.e. the one implementing both satellite and solar tides. We show the final state of the planet-satellite system at 4.5~Gyr in the form of several figures. The results of simulations with different initial planetary spin periods and planetary semi-major axis are shown further below. Whenever the satellite collided with the planet the simulation was stopped and the final obliquity and spin period of the planet were recorded. No subsequent solar tidal evolution was taken into account.\\ \subsection{Endstates} \begin{figure} \resizebox{\hsize}{!}{\includegraphics[angle=-90]{amoon4G.eps}} \caption{Contour plot of the satellite's semi-major axis (in planetary radii) after 4.5~Gyr of tidal evolution. In the central wedge-shaped region the satellite collides with the planet shortly after its formation (Atobe \& Ida, 2007).} \label{amoon4G} \end{figure} In Fig.~\ref{amoon4G} we plot the semi-major axis of the satellite in planetary radii after 4.5~Gyr of tidal evolution as a function of the mass of the satellite and the initial obliquity of the planet. All subsequent figures will be of this type. The colour coding on the right shows the scale. The deep purple region in the middle of the figure, which flares out with increasing satellite mass, is the region where the satellite falls onto the planet (Atobe \& Ida, 2007). In all the figures the collision region will be this colour. This endstate occurs for $\varepsilon_0$ ranging from approximately $60^\circ$ to $120^\circ$. Here extreme seasonal variations in insolation and temperature may occur (e.g. Williams \& Kasting, 1997) and these could be a concern for terrestrial-type life. The widest satellite orbits are obtained for low-obliquity planets and masses between 0.005$m_p$ and 0.0125$m_p$, and for planets with initial obliquity close to 180$^\circ$. Smaller satellites do not raise a high enough bulge on the planet and thus need more time to expand. It should be noted that in all figures, the structure is symmetrical around $\varepsilon_0 = 90^\circ$. The size of the satellite orbit affects the strength of the tidal bulge on the planet and its pole precession frequency. The latter, especially, could be important for habitability when this frequency is close to one of the eigenfrequencies of the solar system. \\ \begin{figure} \resizebox{\hsize}{!}{\includegraphics[angle=-90]{sync4G.eps}} \caption{Colour surface plot of the period ratio of the planet's rotation and the satellite orbit. A ratio of 1 means the system is in the double synchronous state.} \label{sync4G} \end{figure} Something different occurs for massive satellites, which is clearly seen in Fig.~\ref{sync4G}. Here we plot the ratio of the rotation period of the planet to the orbital period of the satellite. For satellites more massive than 0.02~$m_p$ the system is in the double synchronous state. Atobe \& Ida (2007) call this evolution type C. The final period ratio depends steeply on $m_s$ and scales as $m_s^{14}$ (Atobe \& Ida, 2007) and thus the transition from an evolving system to a double synchronous one with increasing $m_s$ is rather abrupt. Applying this to our own system suggests that the habitability of the Earth would be completely different if the Moon were just 50\% more massive. The double synchronous state also explains the decreasing final satellite semi-major axis with increasing satellite mass: the system reaches the double synchronous state and the final orbit is closer to the planet because of the lower initial ratio of angular momentum in the planet's rotation vs. that in the satellite's orbit. \\ \begin{figure} \resizebox{\hsize}{!}{\includegraphics[angle=-90]{prot4G.eps}} \caption{Colour surface plot of the logarithm of the planet's final rotation period in hours.} \label{prot4G} \end{figure} The corresponding rotation period of the planet after 4.5~Gyr of tidal evolution is shown in Fig.~\ref{prot4G}. Here we plot $\log(P_r/1\,{\rm{hr}})$ as a function of satellite mass and initial planet obliquity. Similar to Fig.~\ref{sync4G} above, there is a very steep increase of the rotation period as a function of satellite mass, and the period reaches a maximum for a mass of approximately 0.02$m_p$. Here the rotation period of the planet is several weeks and the satellite semi-major axis sits between 40 and 45~$R_p$. As the satellite mass increases, the final rotation period of the planet is reduced because of the larger fraction of initial angular momentum residing in the satellite orbit compared to the planet's rotation. Even so, in the double synchronous state the rotation period of the planet is always longer than 96~hours, apart from very close to the edge of the central wedge. \\ As we discussed earlier, these long rotation periods affect the climate on the planet and have implications for the origin and sustenance of life. In the double synchronous case the rotation period of the planet could be a substantial fraction of the relaxation time scale and large diurnal temperature variations could occur. Furthermore, in the double synchronous state there are no tides and thus ocean mixing could stop, thereby reducing possible heat transport to the poles. However, the slow rotation rate may increase the Hadley cell to reach the poles, which increases the heat flow to and the temperature at the polar regions. More studies are needed to decide which of these two effects dominates over the other. Lastly, the long rotation period reduces the $J_2$ moment and causes the precession period of the planet to increase. This issue is discussed below.\\ Systems with very light satellites (say $m_s < 0.005$~$m_p$) may negatively affect the habitability because the lack of a tidal bulge raised by the satellite causes the planet to still rotate rapidly, which reduces ocean mixing. This would decrease the heat flow towards the poles, likely causing sustained low temperatures there that reduce habitability.\\ Even though many systems are in the double synchronous state after 4.5~Gyr, the real test to measure if the system has fully evolved is to determine the final obliquity of the planet. From Atobe \& Ida (2007) we know that when the double synchronous state is reached, the planet's obliquity slowly decreases back to 0 if prograde, or increases to 180$^\circ$ if retrograde. Figure~\ref{obl4G} depicts the obliquity of the planet after 4.5~Gyr of tidal evolution. All double synchronous systems are fully evolved and the planet's obliquity is at 0 or 180$^\circ$. The only cases where the obliquity has a different value is when the system is not synchronous or when the satellite has collided with the planet. For planets whose spin is perpendicular to their orbit, the habitable regions are substantially reduced from planets with moderate to high obliquities (Spiegel et al., 2009). \\ \begin{figure} \resizebox{\hsize}{!}{\includegraphics[angle=-90]{obl4G.eps}} \caption{Contour plot of the final obliquity of the planet in degrees.} \label{obl4G} \end{figure} One last quantity that determines the stability of the obliquity, and potentially the habitability, is the precession frequency of the spin pole of the planet. Figure~\ref{prec4G} plots the logarithm of the precession frequency of the spin pole of the planet, in arcsec per year, as a function of satellite mass and initial obliquity. In order to have a rough idea of whether this system will experience a secular spin-orbit resonance, contours corresponding to resonances with the nodal eigenfrequencies $s_6$ and $s_3$ are indicated. Since $s_4 \sim s_3$ (e.g. Brouwer \& van Woerkom, 1950) we did not include it here. We realise that in a system different from our own Solar system, the eigenfrequencies should not be the same, and thus the contours serve an illustrative purpose only. The precession of the planet is slowest when the rotation period is the longest, and some of these systems could have evolved through the $\dot{\psi}=s_6$ and $\dot{\psi}=s_3$ and $\dot{\psi} = s_4$ secular spin orbit resonances.\\ \begin{figure} \resizebox{\hsize}{!}{\includegraphics[angle=-90]{prec4G.eps}} \caption{Contour plot of the logarithm precession frequency, in arcsec per year, of the planet's spin. Contours for resonances with the eigenfrequencies $s_6$ and $s_3$ are indicated.} \label{prec4G} \end{figure} \begin{figure} \resizebox{\hsize}{!}{\includegraphics[angle=-90]{mdist.eps}} \caption{Cumulative distribution of satellite mass generated through a giant impact. See text and Appendix A for details.} \label{mdist} \end{figure} From the above figures it appears there are generally three outcomes: i) the system is still evolving, ii) the system is in the double synchronous state, or iii) the planet has no satellite (initial planetary obliquity near 90$^{\circ}$). The current Earth-Moon system belongs to the first category. We can estimate the probability of being in the first state once we know the mass distribution of the satellites from giant impacts. We have performed this analysis and detailed it in Appendix A and displayed the result in Fig.~\ref{mdist}. The probability for the system to be still evolving is the probability of the system to not be synchronous, and can be evaluated as \begin{equation} P_H = \frac{\sum_H p_{\varepsilon_0}p_m}{\sum_T p_{\varepsilon_0}p_m} \end{equation} where $p_{\varepsilon_0}$ is the probability of the initial obliquity to be $\varepsilon_0$ and $p_m$ is the probability of the satellite to have a mass $m_s$. The numerator sums over all the non-synchronous cases while the bottom sum is for all cases. The probability of the initial spin has the distribution $p_{\varepsilon_0} d\varepsilon_0 = \frac{1}{2}\sin \varepsilon_0 d\varepsilon_0$ (Kokubo \& Ida, 2007). We find that the total probability that the system is still evolving is 85\%.\\ However, what systems yield a state similar to our own i.e. they have $12 < P_r < 48$~h, $\varepsilon_p < 40^\circ$ $(>140^\circ)$ and $0.005 < m_s < 0.02$~$m_p$? From examining Fig.~\ref{obl4G} we can find the initial obliquity that yields a system with the final planetary obliquity and spin period in the specified ranges. We can then repeat the same procedure as above, which yields a probability of 14\% for the system to still be evolving, have $\varepsilon_p<40^\circ$ or $\varepsilon_p>140^\circ$ and $12<P_r<48$~h. This value is somewhat uncertain because there is no unique method to obtain this probability, and thus it only serves as a rough estimate.\\ There are several effects that we have not taken into account, such as the planet's distance from the Sun, the initial rotation period of the planet and the effect of the perturbations of the other planets on the obliquity of the planet. Each of these will be discussed in turn. \subsection{Secular spin-orbit resonance crossing} It is unlikely that extrasolar terrestrial planets will exist in isolation, and recent data confirms this hypothesis (Lissauer et al., 2012). The existence of multiple planets in a given system raises the possibility of the system crossing secular-spin orbit resonances. The chances of that happening depends on the secular architecture of the system. Atobe et al. (2004) studied the effect of the obliquity evolution of terrestrial exoplanets that were perturbed by a giant planet. They concluded that the terrestrial planet's obliquity variations were too large to sustain life if the giant planet was closer than about 5 Hill radii from the terrestrial planet. However, their study was necessarily limited to a few representative cases of planetary configurations because a general study is currently unfeasible. \\ The only planetary system for which we know the secular eigenfrequencies with a decent precision is our own. Ideally we would like to test how a planet-satellite system responds to a secular spin-orbit resonance crossing in a multitude of systems, preferably even in a generalised manner, but that is not possible at this stage. Thus in what follows we shall focus on a fictitious planet-satellite system crossing a secular spin-orbit resonance in our own solar system as a proof of concept, and cautiously use it as a possible outcome for other systems. If the mutual inclinations of the planets in other systems is small, the outcome presented here should be quantitatively similar to what happens in other systems.\\ Figure~\ref{prec4G} showed that it is possible for the planet-satellite system to cross several secular spin-orbit resonances, mostly $\dot{\psi}=s_6$ and $\dot{\psi}=s_3$. The effect of such a resonance crossing is to increase the planet's obliquity. During a secular spin-orbit resonance the obliquity oscillates around its resonant value, $\varepsilon_r$. While in resonance, the maximum libration amplitude of the obliquity is given by (Atobe et al., 2004) \begin{equation} |\Delta \cos \varepsilon_r|_{\rm{max}} = \sqrt{2N_i\sin 2\varepsilon_r}, \end{equation} where $N_i$ is the forced inclination on the planet's orbit corresponding to the frequency $s_i$. If the oscillation range is small, then we can use the approximation $|\Delta \varepsilon_r|_{\rm{max}} = 2\sqrt{N_i\cot \varepsilon_r}$. The increase in obliquity when crossing a resonance is then $\sim 2|\Delta \varepsilon_r|$. Typically $N_i\cot \varepsilon_i = O(10^{-3})$ to $O(10^{-2})$ depending on the resonant obliquity and the forcing, and thus typically $\Delta \varepsilon_r \sim 5^\circ$-10$^\circ$.\\ \begin{figure} \resizebox{\hsize}{!}{\includegraphics[angle=-90]{rescross.eps}} \caption{Obliquity evolution of the planet when passing through the $\dot{\psi}=s_6$ resonance for two different expansion rates of the satellite orbits. The jump in obliquity is consistent with the theory presented in Atobe et al. (2004). The large obliquity ranges near the right at the top panel are caused by the obliquity crossing the resonances $\dot{\psi}=s_3$ and $\dot{\psi}=s_4$. These resonances overlap and the motion is chaotic.} \label{rescross} \end{figure} In Fig.~\ref{rescross} we portray the obliquity of the planet vs time as it goes through the resonance crossing $\dot{\psi}=s_6$. The precession and obliquity of the planet were integrated using the method described in Brasser \& Walsh (2011), and the precession constant was artificially enhanced so as to match the precession rate of the planet just before it crosses the resonance with $s_6$. During the integration the precession constant was linearly decreased in order to mimic the regression of the satellite from the planet. In the top panel the regression was twice as fast as in the bottom panel. One can see that, when crossing the resonance, the obliquity jumps by about 10$^\circ$ in a few million years in the bottom panel while it completes one resonant oscillation in the top panel. In the rightmost part of the top panel the planet hits the resonance $\dot{\psi}=s_3$ and the obliquity oscillates chaotically with large amplitude. For a low-obliquity planet the crossing through the resonance with $s_6$ could have disastrous consequences if it has polar ice caps such as the Earth because the increased insolation at high latitudes would partially melt these ice caps, raise sea levels and potentially wipe out a substantial portion of coastal life. Thus we argue that for terrestrial-type life to develop and be sustained, passage through secular spin-orbit resonances should be avoided.\\ At present the precession rate of the Earth's spin pole is $\dot{\psi} = -50.5$~$''$~yr$^{-1}$. However, a 50\% increase in the lunar mass or initial obliquity of the Earth would have resulted in the Earth-Moon system having passed through the $\dot{\psi} = s_6$ resonance before 4.5~Gyr. \subsection{The effect of initial planetary rotation period} Recent simulations of terrestrial planet formation with a realistic accretion scenario demonstrated that these planets have rotation periods of about twice their minimum value, with $P_{\rm{min}}=2\pi\sqrt{R_p^3/Gm_p}$ and a spread of about $P_{\rm{min}}$ (Kokubo \& Genda, 2010). In order to account for the different initial rotation period of the planet we performed a series of simulations where we set the planet's initial period equal to 3~h and 7~h. Rather than show the full results we decided to only show the regions where the system is doubly synchronous. The results are plotted in Figs~\ref{s43h} and~\ref{s47h}. In the case the initial rotation period is 3~h the minimum satellite mass for which the system is synchronous has now increased beyond 0.03~$m_p$, while it is close to 0.0125~$m_p$ for the case where the initial rotation period was 7~h. Another interesting feature is that the area in the plot where the satellite falls onto the planet also depends on the initial rotation period, and it grows as the initial period increases. The reason for this behaviour is that the tidal evolution increases the obliquity of the planet as the satellite recedes from the planet. At high obliquity the satellite will eventually turn around and fall onto the planet, but this only happens if the inclination between the spin of the planet and the orbit of the satellite $i=\arccos({\bf s}_p\cdot{\bf k}_s)>90^\circ$ part of the time. The planet's obliquity increase is lower for fast-spinning planets, a smaller number of cases will experience retrograde motion and the satellite will not fall onto the planet. \begin{figure} \resizebox{\hsize}{!}{\includegraphics[angle=-90]{sync4G-3h.eps}} \caption{Colour surface plot of the period ratio of the planet's rotation and the satellite orbit. The planet's initial spin period is 3~h. A ratio of 1 means the system is in the double synchronous state.} \label{s43h} \end{figure} \begin{figure} \resizebox{\hsize}{!}{\includegraphics[angle=-90]{sync4G-7h.eps}} \caption{Colour surface plot of the period ratio of the planet's rotation and the satellite orbit. The planet's initial spin period is 7~h. A ratio of 1 means the system is in the double synchronous state.} \label{s47h} \end{figure} \subsection{The effect of initial planetary semi-major axis} The habitable zone around the Sun is thought to reside between 0.7~AU and 1.3~AU (Williams \& Kasting, 1997). The precession time of the planet's spin and the satellite's orbit scales with the planet's semi-major axis as $a_p^{-3}$. This strong dependence means that at the edges of the habitable zone the strength of the solar tides varies by up to a factor of 3 from the value at 1~AU. Thus the regression rate of the planet's spin and the satellite's nodes vary by the same amount, regressing twice as fast at 0.7~AU compared to 1~AU and twice as slow at 1.25~AU compared to 1~AU. Since the final state of the system is dominated by the tidal interaction between the planet and the satellite rather than by the influence of the Sun, we found that the final outcome is very similar to what was presented in the figures above. However, the different precession rates of the planet's spin and the satellite's nodes are noteworthy, in particular in the case where the planet is farther from the Sun because the region inside the $s_6$ contour of Fig.~\ref{4g125} is greatly enhanced. For planets very close to the Sun ($a<0.5$~AU) the tidal forces from the Sun may begin to dominate over those of the satellite for lunar mass satellites. However, at these close solar distances the Hill sphere of the planet becomes comparable to the maximum semi-major axis of the satellite for planets with low obliquity and approximately lunar mass satellites, and other dynamical effects cannot be ignored. Thus the investigation of these cases require a proper N-body treatment. This is beyond the scope of the current study.\\ \begin{figure} \resizebox{\hsize}{!}{\includegraphics[angle=-90]{prec4G125.eps}} \caption{Contour plot of the logarithm precession frequency, in arcsec per year, of the planet's spin for systems with semi-major axis 1.25~AU. Contours for resonances with the eigenfrequencies $s_6$, $s_3$ and $s_2$ are indicated.} \label{4g125} \end{figure} We have run a simulation of the tidal evolution of a planet-satellite pair where the planet has a semi-major axis 1.25~AU. Most of the results are the same as in the previous examples, apart from the precession frequency of the planet's spin pole. Since $\dot{\psi} \propto a^{-3}$, the precession frequency for an identical system at 1~AU should be a factor $\sim 2$ lower. This is depicted in Fig.~\ref{4g125} above, where we show the precession frequency of the planet's spin after 4.5~Gyr of evolution as a function of the satellite mass and initial obliquity. This should be compared directly with Fig.~\ref{prec4G}. As one may see the contours for $s_6$ and $s_3$ occupy a larger area of phase space and in some extreme cases resonance passage with $s_2$ is possible. As we have demonstrated above, resonance with $s_3$ and $s_4$ causes large-amplitude, long-period oscillations in the obliquity and these have drastic consequences for the climate. Thus these resonances should be avoided. \\ \section{Conclusions} In the previous section we have presented the results from our simplified numerical simulations within a well-defined framework. Within this framework the results are robust. We have attempted to place these results in the context of habitability of tidally-evolved terrestrial planet-satellite systems, although the discussion is somewhat speculative due to the large uncertainties in the habitable conditions.\\ We have performed a large sample of numerical simulations of the tidal evolution of an Earth-like planet with a satellite. The satellite's mass and the obliquity of the planet are considered as the two free parameters; the remaining ones are modelled after the current Earth and Moon. In our simplified model, taken from Goldreich (1966), Touma \& Wisdom (1994), Atobe \& Ida (2007) and Correia (2009) the satellite's orbit remains circular. Rather than attempt to determine the final end state of the tidal evolution we ended the simulations when the system reached an age of 4.5~Gyr. This age was determined by requiring that the current Earth-Moon system is reproduced from the initial conditions of Atobe \& Ida (2007). We determined i) which systems are still evolving, ii) which ones are in the double synchronous state accompanied by a perpendicular spin, and iii) which systems have lost their satellites. Systems with obliquities $60^\circ \lesssim \varepsilon_p \lesssim 120^\circ$ lose their satellites (Atobe \& Ida, 2007). We find that after 4.5~Gyr only 85\% of cases, weighted by the satellite mass derived in Appendix A, are still evolving; the rest have either lost their satellite or have reached the double synchronous state. We also discussed habitability as a function of the planet's obliquity and rotation period in each end state, in terms of diurnal/seasonal temperature and ocean mixing and suggest that states ii) and iii) may be less habitable than i). Modelling accretion of a satellite from debris formed by a giant impact, we estimated the probability for an Earth-mass planet to have the end state similar to our Earth-Moon system (12~h $< P_p <$ 48~h and $\varepsilon_0 < 40^\circ$ or $\varepsilon_0 > 140^\circ$), which might be favourable for habitability, amounts to be only 14\%. Elser et al. (2011) conclude that the probability of a terrestrial planet ending up with a heavy satellite ranges from 2\% to 25\% with an average of 13\%. Combining these results suggests that the probability of ending up with a system such as our own is on average only 2\%. \section{Acknowledgements} The authors are deeply grateful to an anonymous reviewer who offered constructive criticism that greatly improved the quality of the paper. Part of this work was performed while RB visited SI and EK at their respective institutes and he is grateful for their hospitality. \section{Appendix A: Satellite mass distribution from a giant impact} We predict the distribution of mass of a planetary satellite accreted from an impact-generated disc. The disc is the result of a collision between a protoplanet and a planetary embryo and the satellite will form from this disc. The results below are based on SPH simulations of giant impacts (Canup et al. 2001; Canup 2004) and N-body simulations of accretion of satellites (Ida et al. 1997; Kokubo et al. 2000). The purpose here is to get a rough distribution but not to pursue detailed fitting with the simulations.\\ Ida et al.(1997) and Kokubo et al.(2000) considered conservation of mass and angular momentum during the accretion: \begin{eqnarray} m_d & \simeq m_{\rm{acc}}+m_{s} & \\ L_{d} & \simeq L_{\rm{acc}}+L_{s} & \simeq m_{\rm{acc}} \sqrt{Gm_{p}R_{p}} + m_{s} \sqrt{Gm_{p}a_{s}}, \end{eqnarray} where $m_{d}$ and $L_{d}$ are mass and orbital angular momentum of the disc, $m_{\rm{acc}}$ and $L_{\rm{acc}}$ are those of the disc materials that are eventually accreted to the host planet, $m_{s}$ and $L_{s}$ are those that are finally incorporated into the satellite, $a_{s}$ is its semi-major axis, and $m_{p}$ and $R_{p}$ are a mass and a physical radius of the planet after the giant impact. Here we neglected the disc materials that escape from the system. From these equations, \begin{equation} \frac{L_{d}}{m_{d}} \simeq \left( 1 - \frac{m_{s}}{m_{d}} \right) \sqrt{Gm_{p}R_{p}} + \frac{m_{s}}{M_{d}} \sqrt{Gm_{p}a_{s}}. \label{eq:disk_L} \end{equation} Canup et al. (2001) compiled the data of previous SPH simulations of giant impacts with $\gamma \equiv m_2/(m_2+m_1) = 0.3$ and ANEOS for an equation of state, where $m_1$ and $m_2$ are target and projectile masses. They found that $m_{d}/m_{p}$ ($m_{p} \simeq m_1+m_2$) and $L_{d}/L_{g}$ are given by a function of only $L_{i}/L_{g}$, where $L_{i}$ is impact angular momentum and $L_{g}$ is that of a parabolic grazing impact. These are given by \begin{equation} \begin{array}{ll} L_{i} & {\displaystyle = \frac{m_1m_2}{m_1 + m_2} R_{i} v_{\rm{esc}}\delta} \\ L_{g} & {\displaystyle = \frac{m_1 m_2}{m_1 + m_2} (R_1 + R_2) v_{\rm{esc}} = \mu \sqrt{2G(m_1+m_2)(R_1 + R_2) } } \\ & {\displaystyle = \sqrt{2} [\gamma^{1/3}+(1-\gamma)^{1/3}]^{1/2} \mu \sqrt{G m_{p} R_{p}}}, \end{array} \label{eq:L_graz} \end{equation} where $R_{i}$ is the impact parameter for a two-body encounter, $R_1$ and $R_2$ are physical radii of the target and the projectile ($R_1 > R_2$), $\mu = m_1 m_2/(m_1+m_2) \simeq \gamma (1-\gamma) m_{p}$, and we assumed that $m_{p} = m_1 + m_2$ and the internal densities of the projectile and the target are the same. In addition from angular momentum conservation we have $\delta = \sqrt{1+v_{\infty}^2/v_{\rm{esc}}^2}$ with $v_{\infty}$ being the speed of the impactor at the planet's Hill radius. \\ Using the compiled data in Canup et al. (2001), we found that the specific angular momentum of the disc is given approximately by \begin{equation} \frac{L_{d}}{m_{d}} \simeq 1.5 \frac{L_{g}}{\mu}. \label{eq:L_specific1} \end{equation} Canup (2004) performed SPH simualtions of giant impacts with $\gamma = 0.11$--0.15 with the M-ANEOS (Melosh 2000) equation of state. The new equation of state results in relatively small $L_{d}/m_{d}$, \begin{equation} \frac{L_{d}}{m_{d}} \simeq 1.2 \frac{L_{g}}{\mu}. \label{eq:L_specific2} \end{equation} Note that Canup et al.(2001) and Canup (2004) did not present the relations (\ref{eq:L_specific1}) and (\ref{eq:L_specific2}). We deduced these relations from their results. Substituting eqs.~(\ref{eq:L_specific1}) and (\ref{eq:L_specific2}) into equation (\ref{eq:disk_L}), we obtain \begin{equation} \left( \sqrt{\frac{a_{s}}{R_{p}}} - 1 \right) m_{s} = \left( \sqrt{2} [\gamma^{1/3}+(1-\gamma)^{1/3}]^{1/2} \alpha - 1 \right) m_{d}, \label{eq:Ms_Md} \end{equation} where $\alpha \simeq 1.5$ for the old ANEOS and $\alpha \simeq 1.2$ for the new ANEOS. The effects of difference in the equation of state are expressed by a slight difference in $\alpha$. \\ From the relations (\ref{eq:L_specific1}) and (\ref{eq:L_specific2}) it appears that an impact-generated disc would be formed from materials of a projectile that do not overlap the target in the line of relative motion. We here follow the conventional low-velocity oblique impact scenario (e.g., Canup et al. 2001) in order to make the discussion simple, although an impact with higher velocity and a steeper angle could also contribute to formation of a large satellite (Reufer et al. 2012). This hypothesis allows us to estimate $m_{d}$ from simple geometrical arguments. Here $\Delta R = R_{imp} - (R_1 - R_2)$ expresses the scale of a part of the projectile forming the disc and $m_{d}$ can be estimated as \begin{eqnarray} m_{d} &\sim& \rho \left( \frac{\Delta R}{2} \right)^3 = \frac{1}{8} \rho R_{p}^3 \left( \frac{\Delta R}{R_1+R_2} \right)^3 \left( \frac{R_1+R_2}{R_{p}} \right)^3 \nonumber \\ &\sim& \frac{1}{8} m_{p} \left( \frac{\Delta R}{R_1+R_2} \right)^3. \label{eq:M_d_estimate0} \end{eqnarray} where we used that $[(R_1+R_2)/R_{p}]^3$ is almost constant for $\gamma \sim 0.1$--0.3. We set \begin{equation} m_{d} = \frac{\beta}{8} m_{p} \left( \frac{\Delta R}{R_1+R_2} \right)^3. \label{eq:M_d_estimate} \end{equation} where $\beta$ is a constant of $O(1)$. From the definition of $\Delta R$, \begin{equation} \frac{\Delta R}{R_1+R_2} = \frac{R_{imp}}{R_1+R_2} - \frac{R_1-R_2}{R_1+R_2} = \frac{L_{i}}{L_{g}\delta} - \xi, \label{eq:DelR_estimate} \end{equation} where \begin{equation} \xi = \frac{(1-\gamma)^{1/3} - \gamma^{1/3}}{(1-\gamma)^{1/3} + \gamma^{1/3}}. \label{eq:xi} \end{equation} which is $\simeq 0.14, 0.28, 0.30$ for $\gamma = 0.3, 0.15, 0.1$, respectively. From eqs.~(\ref{eq:M_d_estimate}) and (\ref{eq:DelR_estimate}) we have \begin{equation} \frac{m_{d}}{m_{p}} \simeq \frac{\beta}{8} \left( \frac{L_{i}}{L_{g}\delta} - \xi \right)^3. \label{eq:Md_Lgraz} \end{equation} If we adopt $\beta \sim 1.2$, this equation reproduces all the results in Canup et al. (2001) ($\gamma = 0.3$, the old ANEOS), Canup \& Asphaug (2001) ($\gamma = 0.1$, Tillotson), and Canup (2004) with ($\gamma = 0.11$-0.15). That is, the scaling relationship in eq.~(\ref{eq:Md_Lgraz}) with projectile mass to total mass ratio ($\gamma$) and a (normalised) impact parameter $L_{i}/L_{g}$ is successful.\\ Now we substitute eq.~(\ref{eq:Md_Lgraz}) into eq.~(\ref{eq:Ms_Md}) to obtain \begin{equation} \frac{m_{s}}{m_{p}} \sim \frac{\beta}{8} \frac{\sqrt{2} [\gamma^{1/3}+(1-\gamma)^{1/3}]^{1/2} \alpha - 1} {\sqrt{a_{s}/R_{p}} - 1} \left( \frac{L_{i}}{L_{g}} - \xi \right)^3. \label{eq:final} \end{equation} Since $[\gamma^{1/3}+(1-\gamma)^{1/3}]^{1/2} \sim 1.2$--1.25 for $\gamma = 0.1$--0.3 and Ida et al.(1997) shows that $a_{s} \sim 3.8 R_{p}$, eq.~(\ref{eq:final}) reads as \begin{equation} \frac{m_{s}}{m_{p}} \sim C \left( \frac{L_{i}}{L_{g}\delta} - \xi \right)^3. \label{eq:final2} \end{equation} with $C \sim 0.17$--0.25 ($\alpha \sim 1.2$--1.5). If one were to set $L_i=(2/5)m_p R_p^2\nu$, where $\nu$ is the planet's rotation rate (see Section 2), together with the first equation in (\ref{eq:L_graz}) one has a relation between the mass of the satellite and the rotation rate of the planet, though it is not a one-on-one relation. For example, we consider the case of $\gamma = 0.13$ ($\xi=0.31$) and $m_{p}=1\,m_\oplus$. In this case, a collision with $L_{i}=1.2L_{EM}$ corresponds to $L_{i}/L_{g}=0.72$ (eq.~\ref{eq:L_graz}). Then, eq.~(\ref{eq:final2}) yields $m_{s} \sim 0.014 m_\oplus \sim 1.1 m_{\rm{Moon}}$, where we used $C \sim 0.2$.\\ We used equation~(\ref{eq:final}) in a Monte-Carlo method to determine the cumulative distribution of satellite masses. We randomly selected $\gamma \in (0.05, 0.5)$, $v_{\infty}$ was taken from a Maxwellian with maximum velocity equal to the planet's escape velocity, and $R_i$ was chosen from $R_i \in (0.2, 1)$ in units of $R_p$ with $R_i^2$ being uniform. The resulting distribution is shown in Fig.~\ref{mdist}. \section{References} Atobe K., Ida S., Ito T., 2004, Icar, 168, 223\\ Atobe K., Ida S., 2007, Icar, 188, 1\\ Barron E.~J., 1983, ESRv, 19, 305\\ Barron E.~J., 1989, Geomorphology 2, 99\\ Bice, K., Marotzke, J., 2002, Paleoceanography 17, 1018\\ Bills B.~G., Ray R.~D., 1999, GeoRL, 26, 3045\\ Bou{\'e} G., Laskar J., 2006, Icar, 185, 312\\ Brasser R., Walsh K.~J., 2011, Icar, 213, 423\\ Brouwer, D., van Woerkom, A. J. J. 1950. Astron. Papers Amer. Ephem. 13, 81-107.\\ Bulirsch, R., Stoer, J. 1966. \ Numerische Mathematik 8, 1-13.\\ Cameron A.~G.~W., 1997, Icar, 126, 126\\ Canup R.~M., Asphaug E., 2001, Natur, 412, 708\\ Canup R.~M., Ward W.~R., Cameron A.~G.~W., 2001, Icar, 150, 288\\ Canup R.~M., 2004, ARA\&A, 42, 441\\ Chambers J.~E., 2001, Icar, 152, 205\\ Correia A.~C.~M., 2009, ApJ, 704, L1\\ Dauphas, N., Pourmand, A., Nature 473, 489\\ Deser, C., Walsh, J.~E., Timlin, M.~S., 2000, J. Climate 13, 617\\ Egbert G.~D., Ray R.~D., 2000, Natur, 405, 775\\ Elser S., Moore B., Stadel J., Morishima R., 2011, Icar, 214, 357\\ Farrell, B. F., 1990, J. Atmos. Sci., 47, 2986\\ Garrett, C., 2003, Nature 422, 477\\ Goldreich P., 1966, RvGSP, 4, 411\\ Hartmann W.~K., Davis D.~R., 1975, Icar, 24, 504\\ Huybers, P, Wunsch, C., 2005, Nature 434, 491\\ Huybers, P., 2011, Nature 480, 229\\ Ida S., Canup R.~M., Stewart G.~R., 1997, Natur, 389, 353\\ Imbrie J., Imbrie J.~Z., 1980, Sci, 207, 943\\ Jutzi, M., Asphaug, E., 2011, Nature 476, 69. \\ Katija, K., Dabiri, J. O., 2009, Nature 460, 624\\ Kinoshita H., Nakai H., 1991, CeMDA, 52, 293\\ Kokubo E., Ida S., Makino J., 2000, Icar, 148, 419\\ Kokubo E., Ida S., 2007, ApJ, 671, 2082\\ Kokubo E., Genda H., 2010, ApJ, 714, L21\\ Lissauer J.~L., 2000, ASPC, 213, 57\\ Lissauer J.~J., et al., 2012, ApJ, 750,112\\ Matsuda, Y. 2000, Planetary meteorology Univ. of Tokyo press, Tokyo, Japan\\ Melosh H.~J., 2000, LPI, 31, 1903\\ Mignard F., 1979, M\&P, 20, 301 \\ Mignard F., 1980, M\&P, 23, 185\\ Milankovi\'{c}, M., 1941. Royal Serbian Sciences 33,633\\ Paillard D., 1998, Natur, 391, 378\\ Reufer A., Meier M., Benz W., Wieler R., 2012, Icarus, in press\\ Spiegel D.~S., Menou K., Scharf C.~A., 2009, ApJ, 691, 596\\ Touma J., Wisdom J., 1994, AJ, 108, 1943\\ Touma J., Wisdom J., 1998, AJ, 115, 1653\\ Webb D.~J., 1982, GeoJI, 70, 261\\ Williams D.~M., Kasting J.~F., 1997, Icar, 129, 254\\ Williams D.~M., Pollard D., 2003, IJAsB, 2, 1\\ Williams G.~P., 1988, ClDy, 3, 45\\ \end{document}
\section{Introduction} Solitons are exceptionally stable wave phenomena that appear in a variety of physical systems. They propagate without spreading due to a fine balance between the dispersion and the non-linearity of the physical system. After their discovery as stable waves in shallow waters by Russell in 1834, solitons have been found in very different non-linear systems. As an example, both bright optical solitons which are peaks in the intensity and dark solitons which are intensity minima have found applications in optical transmission through fibres. In this article we consider soliton waves in Bose-Einstein condensates of neutral atoms. The condensation of a cold gas of atoms permits a description of the many-body problem by a single condensate wave function order parameter, which is in turn described by the non-linear Gross-Pitaevskii wave equation \cite{bog}. The non-linearity in Bose-Einstein condensates arises from the repulsive or attractive interactions of the particles in the condensate. In equivalence with non-linear optics, dark and bright solitons are, respectively, localized dips and peaks in the Bose-Einstein condensate density. The dynamics of solitons in inhomogeneous condensates with slowly varying potentials have been investigated and is adequately described by a particle-like behavior \cite{solitontrain,solitontrain2,4,5,BandADarkSoliton, burger, reinhardt, denschlag}. In this article, we shall further elaborate the analysis of matter wave solitons in 1D-condensates, investigating their scattering properties on localized potentials. In Sec. II, we introduce the Gross-Pitaevskii equation and a convenient mapping to dimensionless coordinates that will be used throughout the paper. In Sec. III, we introduce the bright, the dark, and the composite dark-bright solitons, and we briefly review their properties. In Sec. IV, we show that scattering of bright solitons on a low and wide potential barrier can be well understood as the motion of a classical particle governed by an effective Lagrangian, extracted from the quantum problem, while the scattering on high and narrow barriers can be understood from the single particle quantum problem. In Sec. V we present results for the scattering of a dark-bright soliton on a potential felt only by the bright soliton atoms, and we show to what extent their scattering is transferred to the dark soliton. Sec. VI concludes the paper. \section{Mean field theory of interacting atoms} We treat the Bose-Einstein condensate of atoms at zero temperature in a Hartree mean field theory, which neglects particles outside the condensate and treats short range interactions in the gas by an interaction potential which is proportional to the mean atomic density. This yields the 1D timedependent Gross-Pitaesvkii equation (GPE) \begin{align} i\hbar\frac{\partial \psi(x,t)}{\partial t}&= \bigl(-\frac{\hbar^2}{2m}\frac{\partial ^2}{\partial x^2} +V(x,t)\label{NGPEt}\\\nonumber &+g\abs{ \psi(x,t)} ^2\bigr)\psi(x,t), \end{align} where the order parameter, the macroscopic wave function, is normalized to the total number of atoms \begin{equation} \int \intpartreeNS{r}\abs{\psi(r)} ^2 =N \end{equation} so that $|\psi(x,t)|^2$ represents the particle density. The 1D interaction strength $g$ is a known function of the 3D low energy scattering properties and the transverse confinement of the atoms \cite{bog}.\par In an attractive BEC with $g<0$, (meta-)stable bright soliton solutions with a finite number of particles exist, while in a repulsive BEC with $g>0$, dark solitons may exist in the form of holes in the background density.\par For numerical purposes it is advantageous to rewrite the GPE in a dimensionless form, and for a homogeneous condensate with repulsive interactions, the linear atomic density, $n_0$ and the healing length $\xi=\sqrt{\frac{\hbar^2}{mgn_0}}$ provide natural scale factors. We thus introduce dimensionless quantities for length, energy and time, labelled by a tilde \begin{align} &x=\tilde{x}\xi , \qquad E=\tilde{E}\frac{\hbar ^2}{m\xi ^2}=\tilde{E}n_0 g,\qquad t=\tilde{t}\frac{m\xi ^2}{\hbar}. \label{xEt} \end{align} and obtain the equation \begin{equation} i\frac{\partial \tilde{\psi}}{\partial \tilde{t}}=-\frac{1}{2}\frac{\partial ^2 \tilde{\psi}}{\partial \tilde{x}^2}+\abs{\tilde{\psi}} ^2\tilde{\psi} + \tilde{V}\tilde{\psi}. \label{dimensionlessGPErep} \end{equation}% where the dimensionless wave function \begin{equation} \tilde{\psi}=\frac{\psi}{\sqrt{n_0}}, \label{tildepsi} \end{equation} is normalized to unit bulk density $\abs{\tilde{\psi}}^2 =1$ in the homogeneous part of the condensate. A condensate with attractive interactions has no natural length scale like the healing length, but for convenience in the following, we introduce an arbitrary length unit $l$, which may be chosen to fit a relevant experimental length scale, and we set \begin{equation} x=\tilde{x}l, \qquad E=\tilde{E}\frac{\hbar^2}{ml^2}, \qquad t=\tilde{t}\frac{ml^2}{\hbar}. \end{equation} Inserting this in the GPE, Eq. (\ref{NGPEt}), and introducing the rescaled wave function, $\abs{\tilde{\psi}}^2=\abs{ g}{\frac{ml^2}{\hbar^2}}\abs{\psi}^2$, leads to the dimensionless GPE, \begin{equation} i\frac{\partial \tilde{\psi}}{\partial \tilde{t}}=-\frac{1}{2}\frac{\partial^2 \tilde{\psi}}{\partial \tilde{x}^2}-\abs{\tilde{\psi}}^2\tilde{\psi} +\tilde{V}\tilde{\psi}. \end{equation} The rescaled number of particles in the condensate is \begin{equation} \tilde{N}=\int \intparNS{\tilde{x}}\abs{\tilde{\psi}} ^2 = N\frac{ml\abs{ g}}{\hbar ^2}. \end{equation} Thus for given $N$, the choice of $\tilde{N}$ determines the length scale $l$ and sets the dimensions of the problem. Due to the presence of the non-linear term it is in general necessary to solve the Gross-Pitaesvkii equation by numerical means. The numerical results presented in the following are based on propagation of the one-dimensional, dimensionless and non-homogeneous GPE using the Crank-Nicolson method. For brevity we will omit the tilde above dimensionless quantities throughout. \section{Solitons} \label{solitons} In this section we recall the soliton solutions of the homogeneous Gross-Piaevskii equation. \subsection{Dark solitons} Although the scattering properties of dark solitons are not discussed in this paper we will first introduce the dark soliton. This will serve as a background for the later discussion of the dark-bright soliton.\par Dark solitons are formed in repulsive BECs, governed by the homogeneous GPE, Eq. (\ref{dimensionlessGPErep}), and fulfilling the boundary condition $\lim_{x\rightarrow \pm \infty}\abs{\psi(x)}^2=n_0$, where $n_0$ is the bulk value of the condensate density and also the dimensionless chemical potential. A dark soliton soliton moving with velocity $v$ is then described by \begin{align} &\psi_\mathrm{D}(x,t)=\sqrt{n_0}\Biggl[i\frac{v}{c}+\sqrt{1-\frac{v^2}{c^2}}\tanh\left(\frac{x-vt}{\xi_v}\right)\Biggr]\Exp{-i\mu t/\hbar}, \label{dark} \end{align} where $\xi_v=\xi/\sqrt{1-\frac{v^2}{c^2}}$, with $\xi$ the healing length and $c=\sqrt{\frac{n_0 g}{m}}$ the speed of sound, both unity in our dimensionless units \cite{bog}. The dark soliton has a density dip around $x=vt$, and we will characterize the size of the soliton by the number of atoms missing, $N_\mathrm{D}$, compared to the homogeneous case, \begin{align} N_\mathrm{D}&=\int \intparNS{x} \left(n_0-\abs{\psi_\mathrm{D}} ^2\right)=2n_0\xi\sqrt{1-\frac{v^2}{c^2}}. \label{singelN_D} \end{align} The density depression moves in the opposite direction of the atoms in the condensate, whose motion is given by the gradient of the phase of the condensate wave function. The energy of the dark soliton is given by the difference between the energy of the condensate containing a dark soliton and the energy of the homogeneous condensate, \begin{align} E_\mathrm{D}&=\int \intparNS{x} \Bigl(\frac{\hbar ^2}{2m}\abs{\frac{\partial\psi_\mathrm{D}}{\partial x}} ^2+\frac{g}{2}\left( \abs{\psi_\mathrm{D}} ^2 -n_0\Bigr) ^2 \right)\\\nonumber &=\frac{4}{3}n_0\hbar c \left(1-\frac{v^2}{c^2}\right)^{3/2}. \end{align} The soliton energy decreases with increasing soliton velocity, and for $\frac{v}{c}\ll 1$ one finds \begin{equation} E_\mathrm{D}\simeq \frac{4}{3}n_0\hbar c -\frac{2n_0\hbar}{c}v^2, \end{equation} corresponding to an effective negative mass of $-4n_0\frac{\hbar}{c}$. \subsection{Bright solitons} If the interactions in the BEC are attractive ($g<0$) the homogeneous GPE supports bright soliton solutions, \begin{align} \psi_\mathrm{B}(x,t)=&\sqrt{\frac{2\mu}{g}}\sech\Bigl(\frac{\sqrt{2m\abs{\mu}}}{\hbar}(x-vt)\Bigr)\Exp{i(mvx-\omega t)/\hbar}, \label{brightunits} \end{align} where $\omega =\frac{1}{2}mv^2 +\mu$, $\mu$ being the chemical potential \cite{bog}. The bright soliton propagates with velocity $v$, and contrary to the dark soliton, its shape does not depend on its velocity, but only on the number of atoms in the condensate. In the dimensionless coordinates, the soliton solution, Eq. (\ref{brightunits}), is \begin{align} \psi_\mathrm{B}(x,t)=A\sech \left[ B(x-vt)\right] \Exp{i\left(vx-\omega t\right)}, \label{solution} \end{align} where $A=B=N/2$, the chemical potential of the condensate is $\mu = -\frac{B^2}{2}$, and $\omega =\frac{v^2}{2}-\frac{B^2}{2}$. The energy of the bright soliton, in the absence of an external potential is given by three contributions, \cite{6}: \begin{equation} E=E_\mathrm{kin}^\mathrm{d}+E_\mathrm{kin}^v+E_\mathrm{int}. \end{equation} The first part is the quantum pressure contribution to the kinetic energy \begin{equation} E_\mathrm{kin}^\mathrm{d}=\frac{1}{2}\int \intparNS{x} \abs{ \frac{\partial \abs{\psi}}{\partial x}} ^2 = \frac{N^3}{24}. \label{Edkin} \end{equation} This is present even when the soliton is stationary, while the contribution to kinetic energy from the phase gradient \begin{align} E_\mathrm{kin}^v&=\frac{1}{2}\int \intparNS{x} \abs{ \abs{\psi}\frac{\partial \Exp{i(vx-\omega t)}}{\partial x}} ^2=\frac{1}{2}Nv^2, \end{align} disappears for a stationary soliton.\par Finally, the interactions between the atoms contribute \begin{equation} E_\mathrm{int}=-\frac{1}{2}\int \intparNS{x} \abs{\psi} ^4 =-\frac{N^3}{12} \label{Eint} \end{equation} to the energy of the soliton.\par In total the energy of the bright soliton is \begin{equation} E_\mathrm{B}=\frac{1}{2}Nv^2-\frac{N^3}{24}. \end{equation} \subsection{Dark-Bright solitons} \label{darkbrightsolitons} A dark-bright soliton is a composite object in a two-component condensates consisting of a bright soliton BEC and a density depression in a repulsive BEC. Here, we consider the case where all interactions, internal and between the two components, are repulsive with equal strengths, and in the rescaled coordinates, the coupled Gross-Pitaevskii Equations of the system read \begin{gather} \begin{aligned} &i\frac{\partial \psi_\mathrm{1}(x,t)}{\partial t}=\Bigl(-\frac{1}{2}\frac{\partial ^2}{\partial x^2}+\abs{\psi_\mathrm{1}(x,t)} ^2\\%\nonumber &\phantom{{}i\frac{\partial \psi_\mathrm{1}(x,t)}{\partial t}=\Bigl(} +\abs{\psi_\mathrm{2}(x,t)} ^2\Bigr)\psi_\mathrm{1}(x,t),\\%\nonumber &i\frac{\partial \psi_\mathrm{2}(x,t)}{\partial t}=\Bigl(-\frac{1}{2}\frac{\partial ^2}{\partial x^2}+\abs{\psi_\mathrm{2}(x,t)} ^2\\%\nonumber &\phantom{{}i\frac{\partial \psi_\mathrm{2}(x,t)}{\partial t}=\Bigl(} +\abs{\psi_\mathrm{1}(x,t)} ^2\Bigr)\psi_\mathrm{2}(x,t), \end{aligned} \label{twocomponentBEC} \end{gather} This non-linear, coupled set of differential equations has a dark-bright soliton solution: \begin{align} \nonumber &\psi_\mathrm{D}(x,t)=\Bigl(i\sqrt{\mu}\sin \alpha+\sqrt{\mu}\cos\alpha\tanh\left[\kappa (x-x_0)\right]\Bigr)\Exp{-i\mu t},\\\nonumber &\psi_\mathrm{B}(x,t)=\sqrt{\frac{N_\mathrm{B}\kappa}{2}}\Exp{ix\kappa\tan\alpha}\sech\left[\kappa (x-x_0)\right]\cdot \\ &\phantom{{}\psi_\mathrm{B}(x,t)=}\Exp{-i\left(\frac{\kappa ^2\tan^2\alpha}{2}-\frac{\kappa^2}{2}+\mu\right)t}, \label{darkbrightsolitonsolution} \end{align where \begin{align} \kappa^2+\frac{\kappa N_\mathrm{B}}{2}=\mu\cos^2\alpha, \label{kappaligning} \end{align} and \begin{align} x_0=vt, \label{center} \end{align} with $v=\kappa\tan \alpha$ \cite{Becker, BogA}. The bright soliton can only exists in the repulsive condensate because it is trapped inside the dark soliton. The dark-bright soliton is wider than the single dark soliton, and the larger the number of particles in the bright component, the wider it is due to the repulsive interaction between the two components. The energy of the two-component condensate can be separated into three parts, \begin{equation} E_\mathrm{DB}=E_\mathrm{D} +E_\mathrm{B} +E_\mathrm{int}, \end{equation} where the energy of the dark soliton $E_\mathrm{D}$ is defined by subtracting the energy of the dark background condensate, \begin{gather} \begin{aligned} E_\mathrm{D}&=\int \intparNS{x} \left( \frac{1}{2}\abs{\frac{\partial\psi_\mathrm{D}}{\partial x}} ^2 +\frac{1}{2}\left( \abs{\psi_\mathrm{D}} ^2 -\mu\right) ^2\right)\\ &=\frac{2}{3}\kappa\mu\cos^2\alpha +\frac{2}{3}\frac{\mu^2\cos^4\alpha}{\kappa}, \end{aligned} \end{gather} $E_\mathrm{B}$ is the energy of the bright soliton \begin{gather} \begin{aligned} E_\mathrm{B}&=\int\intparNS{x} \left(\frac{1}{2}\abs{\frac{\partial\psi_\mathrm{B}}{\partial x}} ^2 +\frac{1}{2}\abs{\psi_\mathrm{B}} ^4\right)\\ &= \frac{1}{2}N_\mathrm{B}v^2 +\frac{1}{6}N_\mathrm{B}\kappa\left(N_\mathrm{B} +\kappa\right) , \end{aligned} \end{gather} and $E_\mathrm{int}$ is the energy caused by the interaction between the two components \begin{align} E_\mathrm{int}& =\int \intparNS{x}\abs{\psi_\mathrm{B}} ^2\abs{\psi_\mathrm{D}} ^2\\\nonumber &=N_\mathrm{B}\mu\sin^2\alpha +\frac{1}{3}N_\mathrm{B}\mu\cos^2\alpha . \end{align} The total vector soliton energy can then be written as \citep{BogA} \begin{equation} E_\mathrm{DB}=\frac{4}{3}\kappa ^3 +\frac{1}{2}N_\mathrm{B}\kappa^2 + \frac{1}{2}N_\mathrm{B}v^2 +\mu N_\mathrm{B}. \end{equation} \section{Scattering of Bright solitons} \label{Scatteringofbrightsolitons} Bright solitons are localized objects and in weakly varying potentials where the non-linearity is important, they largely behave as classical particles \cite{shoulder,1,2,3,11,Ching-Hao}, whereas in strongly varying potentials we expect wave diffraction phenomena to enter \cite{Ching-Hao,2,3,11,lee,6,7,weiss,holmer,delta,lang,Parker}. In this Section, we shall investigate the transition between particle and wave behavior of bright solitons scattering on simple square potentials in one dimension. \subsection{Weak Barriers} \label{weakbarriers} As mentioned in the previous section, the chemical potential of a bright soliton is given by $\mu=-\frac{N^2}{8}$, and when the potential is weak compared to the chemical potential, $N^2\gg V_\mathrm{ex}$, we expect the strong interaction potential to hold the atoms together throughout the scattering process.\par \begin{figure*}[htbps] \footnotesize \centering \subfloat[\label{v=0.24}]{\includegraphics[scale=1]{waterfallN1V004b10v024-crop}} \qquad \subfloat[\label{v=0.29}]{\includegraphics[scale=1]{waterfallN1V004b10v029-crop}} \qquad \subfloat[\label{v=0.3}]{\includegraphics[scale=1]{waterfallN1V004b10v03-crop}} \qquad \subfloat[\label{v=0.4}]{\includegraphics[scale=1]{waterfallN1V004b10v04-crop}} \caption{(color online) Time evolution of the condensate density. The external potential, indicated by the dashed lines, is a square barrier with height $V_0=0.04$ and width $b=10$ and the incoming soliton has $N=1$. The threshold velocity, Eq. (\ref{v_crit}), is $v_\mathrm{t}\approx 0.289$. In \protect\subref{v=0.24} the incoming soliton velocity is $v=0.24$, which is well below the threshold velocity, in \protect\subref{v=0.29} the velocity is $v=0.29$, which is just about the threshold velocity and in \protect\subref{v=0.3} and \protect\subref{v=0.4} the velocities are $v=0.3$ and $v=0.4$ respectively, which are above the threshold velocity.} \label{classical} \end{figure*} In figure \ref{classical} the time evolution of the Gross-Pitaevskii wave function is shown for soliton wave packets, incident on a weak square barrier with four different incident velocities. The calculations show that the soliton keeps its sech-shape and is either reflected or transmitted by the barrier like a classical particle.\par In Figs. \ref{v=0.29} and \ref{v=0.3} the soliton is slowed down dramatically and dwells on the potential for a rather long time, before it is eventually reflected or transmitted regaining its initial shape and speed. We define reflection and transmission coefficients as the weight of the wave function on the left and right hand side of the barrier \begin{gather} \begin{aligned} &R=\frac{1}{N}\int_{-L/2}^0 \intpar{x}\abs{ \psi} ^2 \\ &T=\frac{1}{N}\int_0^{L/2} \intpar{x}\abs{ \psi} ^2, \end{aligned} \label{TandRnum} \end{gather} where $L$ is the length of the box used for our calculations, and the barrier is centred at $x=0$. In Fig. \ref{klassisk} transmission and reflection coefficients are shown for two different barrier widths. The figure confirms that the solitons are almost exclusively reflected or transmitted by the wide barrier, while the narrow barrier causes a minor splitting of the mean field wave packet.\par \begin{figure*}[] \footnotesize \subfloat[\label{b=10}]{\includegraphics[scale=0.99]{lagrangeN1V004b10-crop}} \qquad \subfloat[\label{b=2}]{\includegraphics[scale=0.99]{lagrangeN1V004b2-crop}} \caption{(Color online) Transmission and reflection coefficients of a soliton scattering on a weak square barrier for various soliton velocities found by a numerical solution of the GPE (symbols) and by the Lagrangian method (dashed lines). The soliton has $N=1$ and the barrier height is $V_0=0.04$. In \protect\subref{b=10} the barrier has width $b=10$, and is thus wide compared to the soliton width $B^{-1}=2$. In \protect\subref{b=2} the barrier has width $b=2$ which is comparable to the soliton width.} \label{klassisk} \end{figure*} To get a better understanding of the observed phenomena, we assume that the soliton keeps its sech-shape during the scattering process, a fair assumption as seen from the numerical calculations in Fig. \ref{classical}, \begin{equation} \psi(x,t)=A\sech \left[ B(x-x_0)\right] \Exp{i\dot{x}_0\left(x-x_0\right)+i\phi}, \end{equation} and we use the Lagrange approach presented in \cite{1,2,3} to identify equations of motion for the time dependent parameters $A$, $B$, $x_0$ and $\phi$. The effective Lagrangian of the system is \begin{gather} \begin{aligned} \mathcal{L}&=\int_{-\infty}^\infty \intpar{x}\left(\frac{i}{2}\left(\frac{\partial \psi}{\partial t}\psi^\ast - \frac{\partial \psi^\ast}{\partial t}\psi\right)\right) -E(\psi)\\ &=\frac{1}{2}N\dot{x}_0^2-N\dot{\phi}+\frac{1}{6}NB(N-B)-V_{\mathrm{eff}}(x_0), \end{aligned} \label{lagrange} \end{gather} where $V_{\mathrm{eff}}(x_0)$ is \begin{equation} V_{\mathrm{eff}}(x_0)= \int_{-\infty}^\infty \intpar{x}V(x)\abs{\psi(x)}^2. \label{Veffdef} \end{equation} We are now able to find the classical equations of motion for the soliton from the Euler-Lagrange equations \begin{equation} \frac{d}{d t}\frac{\partial \mathcal{L}}{\partial \dot{q}}=\frac{\partial \mathcal{L}}{\partial q}, \qquad q=B,x_0,\phi. \label{euler} \end{equation} Eq.(\ref{euler}) with $q=\phi$ yields $\dot{N}=0$, i.e., the particle number is conserved.\par Eq. (\ref{euler}) with $q=x_0$ yields the equation of motion for the location of the soliton \begin{equation} \ddot{x}_0=-\frac{1}{N}\frac{\partial V_{\mathrm{eff}}(x_0)}{\partial x_0}. \end{equation} This is the equation of motion of a classical particle with mass $N$ and center of mass $x_0$ moving in the effective potential, $V_\mathrm{eff}$, defined by Eq. (\ref{Veffdef}).\par Finally Eq. (\ref{euler}) with $q=B$ yields \begin{equation} N^2-2NB=6\frac{\partial V_{\mathrm{eff}}(x_0)}{\partial B}. \label{Beq} \end{equation} describing a variation of the soliton width as it propagates within the potential barrier.\par For a square barrier, i.e., an external potential of the form \begin{equation} V(x)=\begin{cases} 0 & \text{for $\abs{ x} >\frac{b}{2}$}\\ V_0 & \text{for $\abs{ x} <\frac{b}{2}$}, \end{cases} \label{barrier} \end{equation} the effective potential becomes \begin{align} V_{\mathrm{eff}}(x_0)= \begin{cases} 0 & \text{for $\abs{ x} >\frac{b}{2}$}\\ \frac{V_0N}{2}\bigl(\tanh \left[B\left(\frac{b}{2}-x_0\right)\right]\\ +\tanh\left[B\left(\frac{b}{2}+x_0\right)\right]\bigr) & \text{for $\abs{ x} <\frac{b}{2}$}. \end{cases} \label{Veff} \end{align} The Hamiltonian of the motion is found from the Lagrangian: \begin{equation} H=N\dot{x}_0^2-\mathcal{L}=\frac{1}{2}N\dot{x}_0^2+\frac{1}{6}NB(B-N)+V_\mathrm{eff}(x_0), \end{equation} and the threshold incident energy the soliton needs to cross the barrier is given by \begin{equation} \frac{1}{2}Nv_\mathrm{t}^2-\frac{N^3}{24}=\mathrm{max}_{\abs{ x_0} < \frac{b}{2}}\left\lbrace \frac{1}{6}NB(B-N)+V_\mathrm{eff}(x_0)\right\rbrace . \label{v_crit} \end{equation} The corresponding threshold velocity of the bright soliton is indicated by the dashed lines in Fig. \ref{klassisk}, and reproduces the transition between complete reflection and transmission by the potential.\par We also note that since the barrier is softened by the interparticle interactions, Eq. (\ref{Veff}), it makes sense that the soliton penetrates into the classical forbidden areas as seen in Fig. \ref{classical} \cite{shoulder}. \subsection{Strong Barriers} \label{strongbarriers} Scattering of bright solitons on strong barriers with $V_\mathrm{ex}\gg \mu$ has been studied in the limit of high soliton velocities \cite{2,holmer,6,delta}. In this regime the soliton wave function may coherently split into reflected and transmitted components, and the asymptotic behavior of the transmission and reflection as the soliton velocity increases is the same as for a single particle, described by the linear Schr\"odinger equation. Here, we shall consider scattering of bright solitons on strong potentials with special attention to the behavior at low soliton velocities. Fig. \ref{waterfallV08b05v05} shows the time evolution of a bright soliton incident on a rectangular potential barrier. We observe both a reflected and a transmitted component in the calculation. When the potential barrier is high enough to split the soliton, i.e. $V_0 \gg \mu$, and at the same time wide compared to the soliton width, a part of the soliton is seen to be transiently trapped inside the barrier, see Fig. \ref{zoom}. This trapping inside a potential barrier is due to the wave packet nature of the soliton. The trapped component oscillates and is gradually emitted out of the barrier region. This behavior is not seen when the barrier is narrow compared to the soliton width. \begin{figure*}[htbp] \footnotesize \centering \begin{minipage}[b]{0.47\textwidth} \includegraphics[scale=0.96]{waterfallN1V08b05v05-crop} \end{minipage} \hfill \begin{minipage}[b]{0.47\textwidth} \includegraphics[scale=0.96]{waterfallN1V02b10zoom-crop} \end{minipage} \begin{minipage}[t]{0.47\textwidth} \caption{(Color online) Time evolution of the condensate density. The external potential is a square barrier placed at $x=0$. The potential has height $V_0=0.8$ and width $b=0.5$, and the incoming soliton has $v=0.5$ and $N=1$.} \label{waterfallV08b05v05} \end{minipage} \hfill \begin{minipage}[t]{0.47\textwidth} \caption{(Color online) Time evolution of the condensate wave function. The external potential is a square barrier placed at $x=0$. The potential has $V_0=0.2$ and $b=10$, and the incoming soliton velocity is $v=0.7$. Here we have zoomed in on the time evolution of the wave function $\abs{\psi (x)}$ (not the density), in order to see the details better. } \label{zoom} \end{minipage} \end{figure*} \begin{figure*}[] \footnotesize \subfloat[\label{V=0.6}]{\includegraphics[scale=1]{TRN1V006b2-crop}} \qquad \subfloat[\label{V=0.2}]{\includegraphics[scale=1]{TRN1V002b2-crop}} \qquad \subfloat[\label{V=0.8}]{\includegraphics[scale=1]{TRN1V008b05-crop}} \qquad \subfloat[\label{V=0.4}]{\includegraphics[scale=1]{TRN1V004delta-crop}} \caption{(Color online) Transmission and reflection coefficients of a soliton with $N=1$ on a strong barrier for various incoming soliton velocities. The coefficients has been found by solving the GPE numerically (symbols), by Eq. (\ref{Transmission}) with $E = \frac{1}{2}v^2$ (dashed-dotted lines), and from Eq. (\ref{Transmission}) with $E=\frac{1}{2}v^2-\frac{N^2}{24}$ (dashed lines), see the text. In \protect\subref{V=0.6} the barrier has strength $V_0=0.6$ and width $b=2$, in \protect\subref{V=0.2} the barrier has strength $V_0=0.2$ and strength $b=2$, and in \protect\subref{V=0.8} the barrier has $V_0=0.8$ and $b=0.5$, and thus the same area as in Fig. \protect\subref{V=0.2}. All barriers are higher than the chemical potential, $\abs{ \mu} =\frac{1}{8}$. In \protect\subref{V=0.4} the barrier is a delta barrier of strength $\alpha = 0.4$ which means that the barrier has the same area as the barriers in \ref{V=0.2} and \ref{V=0.8}. For the delta potential the transmission probability reads $T=\Theta (E)\cdot\frac{1}{1+a^2}, \qquad a=\frac{\alpha}{\sqrt{2E}}$.} \label{kvant} \end{figure*} We have solved the GPE numerically for different cases, and Fig. \ref{kvant} summarizes our results (symbols) as functions of the incident velocity. When the external potential is strong compared to the chemical potential of the soliton, we expect the interaction term in the GPE to be less important, and the results in Fig. \ref{kvant} may be understood from a linear approximation based on the Schr\"odinger equation. The solution to the Schr\"odinger equation for a single particle with energy $E$ impacting a square potential barrier as in Eq. (\ref{barrier}) is known, and the transmission coefficient is given by \begin{gather} \begin{aligned} T=&\Theta (E)\times \\ &\begin{cases} \Bigl[1 + \frac{V_0^2}{4E(V_0-E)}\sinh^2 \left(b\sqrt{2(V_0-E)}\right)\Bigr]^{-1} \\ \phantom{{}\Bigl[1 + \frac{V_0^2}{4E(V_0-E)}\sinh^2 \left(b\sqrt{2()}\right)} \text{for $E<V_0$}\\ \Bigl[1 + \frac{V_0^2}{4E(E-V_0)}\sinh^2 \left(b\sqrt{2(E-V_0)}\right)\Bigr]^{-1} \\ \phantom{{}\Bigl[1 + \frac{V_0^2}{4E(E-V_0)}\sinh^2 \left(b\sqrt{2()}\right)} \text{for $E>V_0$}, \end{cases} \end{aligned} \label{Transmission} \end{gather} where the Heaviside step function, $\Theta (E)$, expresses that the energy of the particle scattered must be positive, which is evident in the single particle case, but will be of importance in the following. Ignoring the non-linear interaction term in the GPE is not valid in regions of space where the strong potential is not present. In these regions the soliton behavior is well known, and in particular we know the average energy per particle $\epsilon$: \begin{equation} \epsilon =\frac{E_\mathrm{B}}{N}=\frac{1}{2}v^2-\frac{N^2}{24}. \label{epsilonsol} \end{equation} Because of the attractive interactions between the particles, $\epsilon$ becomes negative when $v < \frac{N}{\sqrt{12}}$ (we have chosen $N=1$ in our calculations which gives $\frac{N}{\sqrt{12}}\approx 0.29$ in our dimensionless units).\par To describe scattering of the soliton in the linear approximation, it makes good sense to replace the energy $E$ in Eq. (\ref{Transmission}) by the average energy per particle in the soliton, $\epsilon$ in Eq. (\ref{epsilonsol}), and in the velocity regime $v \leq \frac{N}{\sqrt{12}}$ the Heaviside stepfunction in Eq. (\ref{Transmission}) suppresses the transmission and predicts full reflection of the soliton. The prediction based on this description is shown in Fig. \ref{kvant} and is in very good agreement with the solution of the GPE. The small transmission probability observed below the velocity threshold may be due, in parts, to our use of wave packets with finite spread, and hence with a momentum distribution exceeding the threshold value. We expect the energy per particle of the reflected and of the transmitted part of the soliton to equal the energy per particle of the initial soliton, $\epsilon_\mathrm{i}$. Since the transmitted (reflected) parts contains a fraction $N_T=T\cdot N$ ($N_R=R\cdot N$) of the atoms, the interaction energy per particle acquires a new value, and we thus expect the velocity of the components to change according to the equation \begin{equation} \epsilon_\mathrm{i} = \frac{1}{2}v_\mathrm{i}^2-\frac{N^2}{24}=\frac{1}{2}v_{T/R}^2-\frac{N^2_{T/R}}{24}. \label{epsiloniandf} \end{equation} Fig. \ref{Efinal} shows examples of the energy per particle in the transmitted and reflected components, determined by the GPE and by the expression (\ref{epsiloniandf}). There is good agreement between the energy per particle in the incoming and in the strongest of the outgoing components, i.e., with the reflected (transmitted) component for low (high) velocities. The poor agreement for the weakest outgoing component is a signature that it is not a pure soliton solution. \begin{figure}[htbp] \footnotesize \centering \begin{minipage}[t]{0.47\textwidth} \includegraphics[scale=0.96]{EfinalEinitialV02b2-crop} \end{minipage} \begin{minipage}[b]{0.47\textwidth} \caption{(Color online) The energy per particle as a function of the incoming soliton velocity. After the scattering the energy per particle is found by numerical simulations of the GPE and Eq. (\ref{epsiloniandf}) (symbols) and the energy per particle of the incoming soliton (solid line) is found by Eq. (\ref{epsilonsol}). } \label{Efinal} \end{minipage} \end{figure} \section{Scattering of Dark-Bright Solitons} \label{Scatteringofdarkbrightsolitons} Dark and dark-bright solitons retain their localized nature when they collide or propagate in slowly varying potentials \cite{bog,10,8,Becker,BogA,Middelkamp,mangevector,darkbrightN}.\par We will here study collisions of dark-bright vector solitons on the rapidly varying square potential, which we choose to only directly affect the bright component. The coupled GP equations for this problem are then given by \begin{gather} \begin{aligned} &i\frac{\partial \psi_\mathrm{D}(x,t)}{\partial t}=\Biggl(-\frac{1}{2}\frac{\partial ^2}{\partial x^2}+\abs{\psi_\mathrm{D}(x,t)} ^2\\ &\phantom{{}i\frac{\partial \psi_\mathrm{D}(x,t)}{\partial t}} +\abs{\psi_\mathrm{B}(x,t)} ^2\Biggr)\psi_\mathrm{D}(x,t),\\ &i\frac{\partial \psi_\mathrm{B}(x,t)}{\partial t}=\Biggl(-\frac{1}{2}\frac{\partial ^2}{\partial x^2}+\abs{\psi_\mathrm{B}(x,t)} ^2\\ &\phantom{{}i\frac{\partial \psi_\mathrm{B}(x,t)}{\partial t}} +\abs{\psi_\mathrm{D}(x,t)} ^2 + V(x)\Biggr)\psi_\mathrm{B}(x,t)\psi_\mathrm{B}(x,t). \end{aligned} \label{twocomponentBECwithV} \end{gather} Because of the rapid spatial variation at the edge of the square potential, we expect the vector soliton to break up into a reflected and a transmitted part as we observed for the single bright soliton. Due to the interaction with the dark component, this system is more complicated, and the linear approximation successfully applied to the bright soliton is not valid here. \begin{figure*} \footnotesize \centering \subfloat[\label{bright}]{\includegraphics[scale=1]{waterfallNB3v05V002b2bright-crop}} \subfloat[\label{dark2}]{\includegraphics[scale=1]{waterfallNB3v05V002b2dark-crop}} \caption{(Color online) Time evolution of the condensate density when the dark-bright soliton scatters on a barrier. \protect\subref{bright} shows the bright component and \protect\subref{dark2} the dark component of the condensate. The incoming soliton has $N_\mathrm{B}=3$ and $\sin\alpha=0.5$, and the barrier, which only directly affects the bright component, has height $V_0=0.2$ and width $b=2$, its location is indicated by the dashed lines. } \label{timedensity} \end{figure*} The results, shown in Fig. \ref{timedensity}, are obtained by numerical simulations of the system (in all numerical calculations $\mu = 1$). The transmission and reflection coefficients, $T_\mathrm{B}$ and $R_\mathrm{B}$, are defined in the same way as for the single bright soliton in Eq. (\ref{TandRnum}). For the dark component, we define the reflection and transmission coefficients as the number of atoms missing with respect to the homogeneous background condensate, on the reflection and transmission side of the barrier, divided by the missing number of atoms $N_D$ in the incident dark component of the vector soliton, \begin{gather} \begin{aligned} &T_\mathrm{D}=\frac{1}{N_\mathrm{D}}\int _0^{x_{\mathrm{sw},T}} \intpar{x} \left(\mu-\abs{ \psi_\mathrm{D}} ^2\right),\\ &R_\mathrm{D}=\frac{1}{N_\mathrm{D}}\int_{x_{\mathrm{sw},R}}^0 \intpar{x} \left(\mu-\abs{ \psi_\mathrm{D}} ^2\right). \end{aligned} \end{gather} The limits $x_{\mathrm{sw},T}$ and $x_{\mathrm{sw},R}$ are determined such that the amplitude variations due to sound waves emitted into the $\psi_\mathrm{D}$-component during the scattering process do not contribute. \begin{figure*} \footnotesize \centering \subfloat[\label{NB3v05}]{\includegraphics[scale=1]{TRNB3v05-crop}} \qquad \subfloat[\label{NB1v05}]{\includegraphics[scale=1]{TRNB1v05-crop}} \caption{Transmission and reflection coefficients of the dark soliton and the bright soliton as a function of the barrier height $V_0$. In \protect\subref{NB3v05} $N_\mathrm{B}=3$ and in \protect\subref{NB1v05} $N_\mathrm{B}=1$. In both cases $\sin\alpha=0.5$ and the barrier width is that of a dark-bright soliton with $\sin\alpha =0$ ($\kappa ^{-1}$, Eq. (\ref{kappaligning})), for $N_\mathrm{B}=3$ that is $b=2$ and for $N_\mathrm{B}=1$ it is $b\approx 1.3$.} \label{TRv05} \end{figure*} We note from Fig. \ref{TRv05} that both the bright and the dark soliton are indeed split by the potential even though only the bright component is directly affected by the barrier. The dark soliton, however, is not reflected to the same degree as the bright soliton, and when the bright soliton reflection reaches unity, the dark soliton transmission converges to a constant which is independent of a further increase in the barrier height, see Fig. \ref{TRv05}. This is also the case for other properties of the system such as the velocities of the outgoing solitons and the amplitudes of the sound waves emitted in the dark component. In Fig. \ref{TRv05} it is also seen that the reflection coefficient and the transmission coefficient of the dark soliton do not add up to one when the barrier height is increased. This we ascribe to sound waves carrying a corresponding surplus of atoms in the background condensate. When $N_\mathrm{B}$ is changed, it alters the soliton scattering properties since the coupling to the bright soliton, mediating the effects of the barrier, is stronger when $N_\mathrm{B}$ is larger. \section{Conclusion} We have investigated scattering of solitons in Bose-Einstein condensates on localized potentials. The homogeneous Gross-Pitaevskii equation supports matter wave solitons, and we have focused on two types of solitons: Bright solitons, which are wave packets without dispersion due to the attractive, non-linear interactions between the atoms in the soliton, and dark-bright solitons in two separate condensate components. In the latter the bright soliton is stabilized by the presence of the dark soliton, a dip in the condensate density. In slowly varying potentials, solitons are known to stay intact and their center of mass motion can be described by a classical equation of motion. We have here considered scattering of solitons on square barriers which vary rapidly at the edges, and where a more quantum mechanical behavior of the soliton is observed. For the bright soliton we found that the scattering properties of the soliton on a localized potential resembles that of classical or quantum mechanical scattering depending on the height and width of the barrier. When the barrier is low compared to the chemical potential of the soliton, the soliton behaves like a robust classical particle, since the barrier cannot overcome the internal interactions of the soliton. The potential appearing in the resulting classical equation of motion describing the soliton center of mass motion is an effective potential modified by the non-linearity. When the barrier is high compared to the chemical potential of the bright soliton, the soliton behaves like a single quantum mechanical particle with a modified relation between its energy and velocity. In this regime, the internal interactions are no longer able to keep the soliton intact and it is split into a reflected and a transmitted part. The scattering properties of a dark-bright soliton on a barrier only affecting the bright component could be expected to resemble that of a single bright soliton scattering on a barrier. But we find that the scattering of a dark-bright soliton is in fact very different. The differences is due to the interaction between the two components of the dark-bright condensate, which enables an indirect interaction between the dark soliton and the external potential. This interaction is strongest when the bright component is large compared to the dark one. Experimental realization of solitons in 1D-condensates is well developed \cite{exppot}, and dark-bright solitons have previously been studied in slowly varying potentials. Thus the prospects of experimental studies along the lines of our calculations seem quite feasible.\par \FloatBarrier
\section{Introduction} \label{Int} Measuring cosmological parameters with observational data plays a crucial role in establishing the modern cosmology. In the past few years, with great progresses on astronomical observations, such as type-Ia supernovae (SN) \cite{union2.1}, cosmic microwave background radiation (CMB) \cite{wmap7} and large scale structure (LSS) \cite{lss}, a standard cosmological model is build up. It is commonly believed that our universe has undergone an inflationary period at the beginning and is accelerating expansion at present. Dark energy, the mysterious source driving the present acceleration of our universe, has been studied widely in the literature since its discovery in 1998 \cite{sn1998}. However, the nature of dark energy, encoded in its equation of state (EoS) parameter $w$, remains controversial. For the other acceleration in the very early universe, the mechanics of inflation can naturally explain the flatness, homogeneity and isotropy of our universe. Inflation stretches the primordial density fluctuations and seeds the presently observed CMB and LSS. Understanding the dynamics of inflation and the origin of late-time acceleration are big challenges in physics and astronomy. With the accumulation of observational data from CMB, LSS and SN observations and the improvements of the data quality, there are a lot of efforts in the literature on constraining the cosmological parameters from various observations \cite{wmap7,Xia:2006cr,constraints}, such as the EoS of dark energy models $w$, the curvature of the universe $\Omega_k$, the total neutrino masses $\sum{m_\nu}$, and those associated with the running of the spectral index and gravitational waves. However, using different input theoretical model can affect constraints of cosmological parameters, since some of cosmological parameters are correlated with each other. Including or ignoring these degeneracies will give significantly different constraints on cosmological parameters. For example, the EoS of dark energy $w$ correlates with the matter energy density $\Omega_m$ and the curvature $\Omega_k$ through the geometrical distance relations. Due to the strong correlation between $w$ and $\Omega_k$, we obtain different constraints on dark energy EoS $w$ in the framework of flat and non-flat universe, respectively \cite{Zhao:2006qg}. The EoS of dark energy also anti-correlates with the total neutrino mass, namely, the larger neutrino mass can be mimicked by the more negative $w$ for the same distance relations \cite{Xia:2006wd,Li:2008vf,Hannestad:2005gj}. In the framework of dynamical dark energy model, the constraint on $\sum{m_\nu}$ is significantly weakened, when comparing with that obtained from the standard $\Lambda$CDM model. Furthermore, the tensor perturbation mode and the dark energy component are correlated, since they mostly affect the large scale (low multipoles) temperature power spectrum of CMB. Thus, including the tensor fluctuations in the analysis will change the constraint of $w$. Moreover, some other cosmological parameters, such as, the initial conditions, inflationary parameters, can also influence the numerical results through the degeneracies. At present, we have a lot of cosmological observations, such as several hundred SN samples, CMB temperature and polarization fluctuation, LSS surveys and so forth. It is possible for us to study the determination of cosmological parameters precisely, especially the parameters correlating with $w$. Thus, in this paper, we study the uncertainties of the constraining cosmological parameters in detail, due to the correlations among them. By employing the Markov Chain Monte Carlo method, we combine the latest observational data, such as seven-year WMAP data (WMAP7), baryon acoustic oscillation (BAO) and the ``Union2.1'' compilation SN sample to constrain cosmological parameters in different input cosmological models, such as the non-flat universe, the massive neutrino mass models, and so on. We pay particular attention on the degeneracies among these parameters. Furthermore, we also generate the mock data for the future observations, such as PLANCK \cite{planck} and Euclid \cite{Euclid,Euclid-official} projects, as well as future SN observations \cite{futureSN} to study their capabilities on breaking these degeneracies. The structure of the paper is as follows: in section \ref{Method} we describe the method and the current and future observational data sets we use; Section \ref{results} contains our main numerical results on the cosmological parameters, and the last section is the summary. \section{Method and Data} \label{Method} \subsection{Method} In order to get constraints in different cosmological models, we employ the modified version of the MCMC package {\tt CosmoMC} \cite{CosmoMC} and perform the global fitting analysis with the current and future observational data. We parameterize the cosmology with the following parameters: \begin{equation} \label{parameter} {\bf P} \equiv (\omega_{b}, \omega_{c}, \Omega_k, \Theta_{s}, \tau, w_{0}, w_{a}, f_{\nu}, n_{s}, A_{s}, \alpha_s, r)~, \end{equation} where $\omega_{b}\equiv\Omega_{b}h^{2}$ and $\omega_{c}\equiv\Omega_{c}h^{2}$, in which $\Omega_{b}$ and $\Omega_{c}$ are the physical baryon and cold dark matter densities relative to the critical density, $\Omega_k$ is the spatial curvature and satisfies $\Omega_k+\Omega_b+\Omega_c+\Omega_{{\mathrm{de}}}=1$, $\Theta_{s}$ is the ratio (multiplied by 100) of the sound horizon to the angular diameter distance at decoupling, $\tau$ is the optical depth to re-ionization, $f_{\nu}$ is the dark matter neutrino fraction at present\cite{Lesgourgues:2006nd}, namely, \begin{equation} f_{\nu}\equiv\frac{\rho_{\nu}}{\rho_{m}}=\frac{\Sigma m_{\nu}}{93.14~\mathrm{eV}~\Omega_mh^2}~. \end{equation} The primordial scalar power spectrum $\mathcal{P}_{s}(k)$ is parameterized as \cite{Ps}: \begin{eqnarray} \ln\mathcal{P}_{s}(k)=\ln A_s(k_{s0})+(n_s(k_{s0})-1)\ln\left(\frac{k}{k_{s0}}\right)+\frac{\alpha_s}{2} \left[\ln\left(\frac{k}{k_{s0}}\right)\right]^2 \end{eqnarray} where $A_s$ is defined as the amplitude of initial power spectrum, $n_s$ measures the spectral index, $\alpha_{s}$ is the running of the scalar spectral index and $r$ is the tensor to scalar ratio of the primordial spectrum. For the pivot scale we set $k_{s0}=0.05\,$Mpc$^{-1}$ which is the default value of {\tt CosmoMC} package. $w_0$ and $w_a$ are the parameters of dark energy, and the EoS is parameterized as \cite{Linderpara}: \begin{equation} \label{EOS} w_{\mathrm{de}}(a) = w_{0} + w_{a}(1-a)~, \end{equation} (noted as ``CPL''), where $a=1/(1+z)$ is the scale factor and it covers $\Lambda$CDM model by taking $w_0=-1$ and $w_a=0$. As we know, for time evolving dark energy model, it is crucial to include dark energy perturbations in the global fitting strategy to constrain cosmological parameters \cite{WMAP3GF,LewisPert,XiaPert}. In this paper we use the method provided in refs.\cite{XiaPert,ZhaoPert} to treat the dark energy perturbations consistently in the whole parameter space in the numerical calculations. \subsection{Observational Data} \subsubsection{Current Data} In our analysis, we consider the following cosmological probes: i) power spectra of CMB temperature and polarization anisotropies; ii) the baryon acoustic oscillation in the galaxy power spectra; iii) measurement of the current Hubble constant; iv) luminosity distances of type Ia supernovae. To incorporate the WMAP7 CMB temperature and polarization power spectra, we use the routines for computing the likelihood supplied by the WMAP team \cite{wmaplike}. The WMAP7 polarization data are composed of TE/EE/BB power spectra on large scales ($2 \leq\ell\leq 23$) and TE power spectra on small scales ($24 \leq\ell\leq 800$), while the WMAP7 temperature data includes the CMB anisotropies on scales $2 \leq\ell\leq 1200$. Baryon Acoustic Oscillations provides an efficient method for measuring the expansion history by using features in the clustering of galaxies within large scale surveys as a ruler with which to measure the distance-redshift relation. The BAO aries because that the coupling of baryons and photons by Thomson scattering in the early universe allows acoustic oscillations in the plasma at early time. After photons decoupling, the acoustic waves can propagate a certain distance, which becomes a characteristic comoving scale and remains in the distribution of galaxies. It provides a particularly roburst quantity to measure\cite{Peebles:1970ag,Sunyaev:1970eu,Eisenstein:1997ik,Meiksin:1998ra,Eisenstein:2006nj}. BAO has been detected in the current galaxy redshift survey data from the SDSS DR7 \cite{lss}. It measures not only the angular diameter distance, $D_A(z)$, but also the expansion rate of the universe, $H(z)$, which is powerful for studying dark energy \cite{task}. Since the current BAO data are not accurate enough for extracting the information of $D_A(z)$ and $H(z)$ separately \cite{okumura}, one can only determine an effective distance \cite{baosdss}: $D_v(z)=[(1+z)^2D_A^2(z)cz/H(z)]^{1/3}$. In this paper we use the BAO measurement given by the SDSS-II survey \cite{bao-SDSSII}. In our analysis, we add a Gaussian prior on the current Hubble constant given by ref. \cite{h0}; $H_0 = 74.2 \pm 3.6$ km\,s${}^{-1}$\,Mpc${}^{-1}$ (68\% C.L.). The quoted error includes both statistical and systematic errors. This measurement of $H_0$ is obtained from the magnitude-redshift relation of 240 low-z Type Ia supernovae at $z < 0.1$ by the Near Infrared Camera and Multi-Object Spectrometer (NICMOS) Camera 2 of the Hubble Space Telescope (HST). This is a significant improvement over the previous prior, $H_0 = 72 \pm 8$ km\,s${}^{-1}$\,Mpc${}^{-1}$, which is from the Hubble Key project final result. In addition, we impose a weak top-hat prior on the Hubble parameter: $H_0 \in [40, 100]$ km\,s${}^{-1}$\,Mpc${}^{-1}$. Finally, we include data from Type Ia supernovae, which consists of luminosity distance measurements as a function of redshift, $D_L(z)$. In this paper we use the latest SN data sets from the Supernova Cosmology Project, ``Union Compilation 2.1'', which consists of 580 samples and spans the redshift range $0 \leq z\leq1.55$ \cite{union2.1}. This data set also provides the covariance matrix of data with and without systematic errors. In order to be conservative, we use the covariance matrix with systematic errors. When calculating the likelihood from SN, we marginalize over the absolute magnitude M, which is a nuisance parameter, as done in refs. \cite{SNMethod}. \subsubsection{Future Measurements} Since the current observations can not give the conclusive conclusion, we also use the simulated data from future observations of Planck and Euclid, as well as the SN sample. For the CMB simulation we consider a simple full-sky ($f_{\rm sky} = 1$) simulation at Planck-like sensitivity. We neglect foregrounds and assume the isotropic noise with variance $N^{\rm TT}_\ell = N^{\rm EE}_\ell / 2 = N^{\rm BB}_\ell / 2 = 3\times10^{-4}\mu K^2$ and a symmetric Gaussian beam of 7 arcminutes full-width half-maximum (FWHM). We use the simulated $C_\ell$ up to $\ell = 2500$ for temperature \footnote{ We also use a more conservative choice $\ell_{\rm max}=1500$ to simulate the mock temperature power spectrum and find that the obtained constraints on cosmological parameters are almost unchanged.} and $\ell = 1500$ for polarization. For the future LSS survey, we simply consider the Euclid-like survey which will measure $\sim 10^8$ galaxies with the redshifts $z< 2$. In the measurements of large scale matter power spectrum of galaxies there are generally two statistical errors: sample variance and shot noise. The uncertainty due to statistical effects, averaged over a radial bin $\Delta k$ in Fourier space, is \cite{lsserror}: \begin{equation} \left(\frac{\sigma_P}{P}\right)^2=2\times\frac{(2\pi)^3}{V_{\rm survey}}\times\frac{1}{4\pi k^2\Delta k}\left(1+\frac{1}{\bar{n}P}\right)^2~. \end{equation} The initial factor of 2 is due to the real property of the density field, $V_{\rm survey}$ is the survey volume, and $\bar{n}$ is the mean galaxy density. In our simulations for simplicity and to be conservative, we use only the linear matter power spectrum up to $k_{\rm max} = 0.1h$\,Mpc${}^{-1}$. For the future supernovae observation, we simulate 4000 SN samples distributed in 17 bins from $z=0.1$ to $z=1.7$. We also include $300$ low-z SN from the Nearby Supernova Factory \cite{WoodVasey:2004pj} to improve the constraints. For all the supernovae samples, we assume the variance $\sigma=0.15$ in each redshift bin. \section{Numerical Results}\label{results} \subsection{Current Data} In this section, we present the numerical results from current observational data sets with different input theoretical models. Here, we mainly focus on the EoS of dark energy, the primordial power spectrum index, and the total neutrino mass. \subsubsection{Dark Energy}\label{sec-DE} \begin{table TABLE I. $1\,\sigma$ constraints on some cosmological parameters from Union2.1+WMAP7+BAO+HST with different input theoretical models. \begin{center \begin{tabular}{c|c|c|c} \hline\hline models &$w_0$&$w_a$ &$\Delta\chi^2$\\ \hline $CPL$&$-0.94\pm 0.18$&$-0.40\pm 0.78$&$-$ \\ \hline $CPL\,\oplus\,\Omega_k$&$-0.93\pm 0.19$&$-0.47^{+1.00}_{-0.97}$&$-0.2$ \\ \hline $CPL\,\oplus\,m_{\nu}$&$-0.92\pm 0.20$&$-0.86\pm 1.00$&$\sim 0$ \\ \hline $CPL\,\oplus\,r$&$-0.97\pm 0.18$&$-0.11^{+0.76}_{-0.75}$&$\sim 0$\\ \hline\hline \end{tabular} \end{center} \end{table} To study the correlations between the EoS of dark energy and other cosmological parameters, such as the curvature of universe, the total neutrino mass, and tensor perturbation mode, we perform several global analyses with different input cosmological models, for example, the flat universe with massless neutrino (noted as ``$CPL$''), the non-flat universe with massless neutrino (noted as ``$CPL\,\oplus\,\Omega_k$''), the flat universe with massive neutrino model (noted as ``$CPL \,\oplus\, m_{\nu}$'') and the flat universe with tensor perturbation (noted as ``$CPL \oplus r$''). We summarize the constraints on $w_0$ and $w_a$ in table I. \begin{figure}[t] \begin{center} \includegraphics[scale=0.6]{w0wa_2D.eps} \caption{2-dimensional cross correlation between $w_0$ and $w_a$ in different input cosmological models: the black solid lines are given by the standard $CPL$ model, the red dashed, blue dash-dot and green dotted lines are given by $CPL\,\oplus\,\Omega_k$, $CPL\,\oplus\,M_{\nu}$, and $CPL\,\oplus\,r$ models, respectively.\label{w0wa}} \end{center} \end{figure} In figure \ref{w0wa}, we plot the 2-dimensional constraints of $w_0$ and $w_a$ with different input cosmological models. The upper left plot is obtained from the standard ``$CPL$'' model. The 68\% C.L. constraints are $w_0=-0.94\pm 0.18$ and $w_a=-0.40\pm 0.78$, which is consistent with previous works \cite{wmap7}. However, if we choose a different input cosmological model, the constraints will change obviously. In the upper right panel of figure \ref{w0wa}, we compare the constraints on $w_0$ and $w_a$ from the flat (black solid lines) and non-flat (red dashed lines) universe cases, respectively. Apparently, the model with non-zero $\Omega_k$ give weaker constraints on dark energy EoS parameters and the median values are shifted, namely, the constraints are $w_0=-0.93\pm 0.19$ and $w_a=-0.47^{+1.00}_{-0.97}$ at the 68\% confidence level. We know that the dark energy and the curvature are correlated via the distance relation. The effect on the distance of dark energy EoS can be mimicked by the non-zero $\Omega_k$. Therefore, the degeneracy between them should be very strong. In our analysis, we find that the correlation coefficients are $-0.32$ and $0.58$ between $\Omega_k$ and $w_0$ or $w_a$, respectively. Then, we consider the effect of tensor perturbation mode on the constraints of dark energy EoS. In the lower left panel of figure \ref{w0wa}, we present the constraint on the EoS parameters of dark energy from the models including tensor perturbation mode (green dotted lines). The tensor perturbation mode mainly contributes the CMB temperature anisotropies on the very large scales. Meanwhile, the dark energy component also affects the low multipoles temperature power spectrum of CMB through the late-time integrated Sachs-Wolfe (ISW) effect \cite{isw}. When including the tensor perturbation mode, we obtain the $1\,\sigma$ constraints: $w_0=-0.97\pm 0.18$ and $w_a=-0.11^{+0.76}_{-0.75}$. The constraints do not change significantly, considering the large error bars. Finally, we investigate the degeneracy between the EoS of dark energy and the massive neutrino. In the lower right panel of figure \ref{w0wa}, we show the constraint on $w_0$ and $w_a$ when considering the massive neutrinos (blue dash-dot lines). In this case, the constraints on the EoS parameters of dark energy are obviously weakened and the median values are shifted: $w_0=-0.92\pm 0.20$ and $w_a=-0.86\pm1.00$ (68\% C.L.). This results in a strong anti-correlation between $\sum{m_\nu}$ and the EoS parameters of dark energy, which mainly comes from the geometric feature of our universe \cite{Hannestad:2005gj}. In addition, dynamical dark energy will modify the time evolving potential wells which affect CMB power spectra through the late time ISW effect. Dynamical dark energy can leave imprints on CMB, LSS power spectra, and Hubble diagram, nonetheless these features can be mimicked by cosmic neutrino to some extent \cite{copeland}. In our analysis, we find that the correlation coefficients are $0.08$ and $-0.37$ between $\sum{m_\nu}$ and $w_0$ or $w_a$, respectively. \subsubsection{Inflationary parameters} \label{sec_IF} Currently, the cosmological observational data are in good agreement with a Gaussian, adiabatic, and scale-invariant primordial spectrum, which are consistent with single-field slow-roll inflation predictions. Measuring the spectral index $n_s$ of primordial power spectrum is important for the studies of inflation paradigm. In the literature, the constraint on $n_s$ is usually obtained by fitting the basic 6-parameter $\Lambda$CDM model. However, the constraints can be changed by the correlations between $n_s$ and other cosmological parameters, such as the running of spectral index $\alpha_s$, defined as $\alpha_s=d n_s/d\ln k$, the tensor perturbation mode, as well as the EoS parameters of dark energy. To investigate these correlations, we constrain $n_s$ in several different input cosmological models: the standard $\Lambda$CDM model which assumes a pure scale invariant power-law power spectrum (noted as $\Lambda$CDM), the standard $\Lambda$CDM model with running spectral index (noted as ``$\Lambda$CDM$\,\oplus\,\alpha_s$''), the standard $\Lambda$CDM model with tensor perturbation mode (noted as ``$\Lambda$CDM$\,\oplus\,r$''), the standard $\Lambda$CDM model with running spectral index and tensor perturbation mode (noted as ``$\Lambda$CDM$\,\oplus\,\alpha_s\,\oplus\,r$''), the standard $\Lambda$CDM model with running spectral index and massive neutrinos (noted as ``$\Lambda$CDM$\,\oplus\,\alpha_s\,\oplus\,m_{\nu}$''), the standard $\Lambda$CDM model with tensor perturbation mode and massive neutrinos (noted as ``$\Lambda$CDM$\,\oplus\,r\,\oplus\,m_{\nu}$''). Furthermore, we have studied the correlations between the dynamical dark energy and inflationary parameters (noted as ``$CPL$'', ``$CPL\,\oplus\,\alpha_s$'' and ``$CPL\,\oplus\,r$'' models, respectively). We list the numerical results in table II and plot the 1-dimensional distributions of $n_s$ when using different input cosmological models. \begin{table TABLE II. $1\,\sigma$ constraints on the Inflationary parameters $n_s$, $\alpha_s$, and $r$ from Union2.1+WMAP7+BAO+HST. For the weakly constrained parameters, we quote the $95\%$ upper limits instead. \begin{center \begin{tabular}{c|c|c|c|c} \hline\hline &$n_s$ &$\alpha_s$&$r$&$\Delta\chi^2$ \\ \hline $\Lambda$CDM&$0.969\pm0.011$&$-$&$-$&$-$\\ \hline $\Lambda$CDM$\,\oplus\,\alpha_s$&$0.951\pm0.020$&$-0.018\pm0.016$&$-$&$-2.0$ \\ \hline $\Lambda$CDM$\,\oplus\,r$&$0.974\pm0.012$&$-$&$<0.15$&$-0.5$ \\ \hline $\Lambda$CDM$\,\oplus\,\alpha_s\,\oplus\,r$&$0.946^{+0.020}_{-0.021}$&$-0.038\pm0.022$&$<0.37$&$-2.2$ \\ \hline $\Lambda$CDM$\,\oplus\,m_{\nu}\,\oplus\,\alpha_s$&$0.960^{+0.021}_{-0.022}$&$-0.012\pm0.017$&$-$&$-1.6$ \\ \hline $\Lambda$CDM$\,\oplus\,m_{\nu}\,\oplus\,r$&$0.960^{+0.024}_{-0.023}$&$-$&$<0.39$&$-1.6$ \\ \hline $CPL$&$0.966\pm0.013$&$-$&$-$&$-0.3$ \\ \hline $CPL\,\oplus\,\alpha_s$&$0.930^{+0.025}_{-0.026}$&$-0.029^{+0.019}_{-0.020}$&$-$&$-2.7$ \\ \hline $CPL\,\oplus\,r$&$0.979^{+0.016}_{-0.018}$&$-$&$<0.20$&$-0.2$ \\ \hline\hline \end{tabular} \end{center} \end{table} \begin{figure}[t] \begin{center} \includegraphics[scale=0.5]{ns.eps} \caption{1-dimensional probability distribution of the spectral index $n_s$ in different input cosmological models. The black solid line is given by the standard $\Lambda$CDM model, while the red dotted, blue dash-dotted line, green dashed are obtained from the $\Lambda$CDM$\,\oplus\,\alpha_s$, $\Lambda$CDM$\,\oplus\,r$ and $\Lambda$CDM$\,\oplus\,m_{\nu}$ models, respectively. For comparison, we also show the distribution from the standard $CPL$ model (purple dash-dot-dot line).\label{ns}} \end{center} \end{figure} \begin{figure}[t] \begin{center} \includegraphics[scale=0.6]{nsas_2D.eps} \caption{2-dimensional cross correlation between $n_s$ and $\alpha_s$ in different input cosmological models: The black solid lines are given by the standard $\Lambda$CDM model, the red dotted lines, blue dash-dot lines and green dashed lines are given by $\Lambda$CDM$\,\oplus\,\alpha_s\,\oplus\,r$, $\Lambda$CDM$\,\oplus\,m_{\nu}\,\oplus\,\alpha_s$ and $CPL\,\oplus\,\alpha_s$ models, respectively.\label{ns-as}} \end{center} \end{figure} In the standard 6-parameter $\Lambda$CDM model, we obtain the $1\,\sigma$ constraint on $n_s$ is $n_s=0.969\pm0.011$, which is consistent with previous works \cite{wmap7}. When considering the running of spectral index, the constraint of $n_s$ is significantly weakened, $n_s=0.951\pm0.020$ (68\% C.L.), while the 68\% constraint of $\alpha_s$ is $\alpha_s=-0.018\pm0.016$. The change of the constraint on $n_s$ is about $1.6\,\sigma$. We also obtain that the correlation coefficient between $n_s$ and $\alpha_s$ is $0.83$, which implies the strong correlation between them. When including tensor perturbation mode, the constraint on $n_s$ slightly changes: $n_s=0.974\pm0.012$ at the 68\% confidence level. When varying both $\alpha_s$ and $r$ simultaneously, the $1\,\sigma$ error bars of Inflationary parameters are significantly enlarged: $n_s=0.946^{+0.020}_{-0.021}$ and $\alpha_s=-0.038\pm0.022$. The correlation coefficients among $n_s$, $\alpha_s$ and $r$ are $0.70$, $0.35$ and $-0.60$, respectively. In figure \ref{ns-as} we plot the 2-dimensional constraints on the panel ($n_s$,$\alpha_s$) in the upper two panels with or without considering the tensor perturbation, respectively. The correlation between the early universe inflation and late time accelerating expansion can also affect the constraints on inflationary parameters. In time-evolving dark energy model ($CPL$), we obtain the 68\% C.L. constraint $n_s=0.966\pm0.013$, which is similar with that in the $\Lambda$CDM framework. However, when considering the running of spectral index, the constraints of $n_s$ and $\alpha_s$ are obviously weakened, namely, $n_s=0.930^{+0.025}_{-0.026}$, and $\alpha=-0.029^{+0.019}_{-0.020}$ at the 68\% confidence level. More importantly, their mean values are shifted significantly, when comparing them from the standard $\Lambda$CDM model. In the ``$CPL\,\oplus\,r$'' model, we obtain the constraints $n_s=0.979^{+0.016}_{-0.018}$ ($1\,\sigma$) and $r<0.20$ ($2\,\sigma$). Their error bars become larger which is clearly shown in the lower right panel of figure \ref{ns-as}, due to the correlation between the dark energy and the tensor perturbation we discuss before. We also study the correlations between the massive neutrino and $n_s$, $\alpha_s$, $r$. With the massive neutrino, we obtain the constraints $n_s=0.960^{+0.024}_{-0.023}$ ($1\,\sigma$) and $r<0.39$ ($2\,\sigma$) in the ``$\Lambda$CDM$\,\oplus\,m_{\nu}\,\oplus\,r$'' model, whose error bars are much larger than those in the massless neutrino case. \subsubsection{Neutrino mass} \label{sec-NM} \begin{table TABLE III. The 95\% confidence level upper limits on the total mass of neutrinos from Union2.1+WMAP7+BAO+HST. \begin{center \begin{tabular}{c|c|c} \hline\hline &$\Sigma m_{\nu}$&$\Delta\chi^2$ \\ \hline $LCDM\oplus m_{\nu}$&$<0.45\,$eV &$-$\\ \hline $LCDM\oplus m_{\nu}\oplus r$&$<0.55\,$eV &$-1.1$\\ \hline $LCDM\oplus m_{\nu}\oplus \alpha_s$&$<0.42\,$eV &$-1.1$\\ \hline $CPL\oplus m_{\nu}$&$<0.81\,$eV &$\sim 0$\\ \hline\hline \end{tabular} \end{center} \end{table} \begin{figure}[t] \begin{center} \includegraphics[scale=0.5]{mu.eps} \caption{1-dimensional probability distributions of the neutrino mass in different input cosmological models. The black solid line is given by the standard $\Lambda$CDM model, while the blue dotted, red dashed and the green dash-dot lines are given by the $\Lambda$CDM$\,\oplus\,m_{\nu}\,\oplus\,\alpha_s$, $\Lambda$CDM$\,\oplus\,m_{\nu}\,\oplus\,r$ and $CPL\,\oplus\,m_{\nu}$ models, respectively.\label{mu}} \end{center} \end{figure} Neutrino, as a hot dark matter candidate, can be constrained from the free streaming modification of the transfer function of the matter power spectrum. Thus, the constraints on neutrino mass are correlated with the shape of primordial power spectrum. Since the free streaming effects can be compensated by the modification of the primordial spectrum, the constraints on total neutrino mass can be influenced by the inflationary parameters, as we discuss above. Moreover, the behaviour of dark energy EoS can also affect the constraint of the total neutrino mass. In table III and figure \ref{mu} we show the numerical constraints on the massive neutrino from different input models: the standard $\Lambda CDM$ model (noted as ``$\Lambda$CDM$\,\oplus\,m_{\nu}$''), the standard $\Lambda$CDM with tensor perturbation mode (noted as ``$\Lambda$CDM$\,\oplus\,m_{\nu}\,\oplus\,r$''), the standard $\Lambda$CDM with running of spectral index (noted as "$\Lambda$CDM$\,\oplus\,m_{\nu}\,\oplus\,\alpha_s$''), and the $CPL$ dynamical dark energy model (noted as ``$CPL\,\oplus\,m_{\nu}$''). In the framework of standard $\Lambda$CDM model, the $2\,\sigma$ upper limit on the total neutrino mass is $\sum m_{\nu} < 0.45\,$eV. When including the tensor perturbation mode, the constraint is slightly weakened, namely, $\sum m_{\nu} < 0.55\,$eV at the 95\% confidence level. As we know, both the negative $\alpha_s$ and the massive neutrino lead to a damped power on small scales. The effect of massive neutrinos may be compensated by a non-vanishing running of the primordial spectrum. On the other hand, the running is negative, this will lead to even more stringent constraints on the neutrino mass compared with fittings in the constant scalar spectral index cosmology \cite{fnurun}. In practice, we find the 95\% upper limit on the total neutrino mass becomes smaller, when varying the running of the spectral index, $\sum m_{\nu} < 0.42\,$eV. The massive neutrino is strongly correlated with the EoS of dark energy. Therefore, the behaviour of dark energy model, especially the dynamical dark energy model, will significantly enlarge the constraint of total neutrino mass. In the ``$CPL\,\oplus\,m_{\nu}$'' model, we obtain the $2\,\sigma$ upper limit on the massive neutrino is $\sum m_{\nu} < 0.81\,$eV, as shown in figure \ref{mu}. \subsection{Future Data} From the results presented above, we see that there are some strong degeneracies among cosmological parameters, such as $w$ and $\sum{m_\nu}$, $n_s$ and $\alpha_s$. Due to the precision of current observations, these degeneracies can not be broken and will weaken the constraints of cosmological parameters. Therefore, it is worthwhile discussing whether future observations could give more stringent constraints on the cosmological parameters and break these degeneracies efficiently. For this purpose, we have performed a further analysis and we have chosen the fiducial model in perfect agreement with current data: $\Omega_bh^2=0.0227$, $\Omega_{c}h^2=0.114$, $\vartheta=1.041$, $\tau=0.0865$, $w_0=-1$, $w_a=0$, $n_s=1$, $A_s=2.85\times 10^{-9}$ at $k_\ast=0.05\,$Mpc${}^{-1}$. \subsubsection{$\Omega_k$ and $w(z)$} \begin{table TABLE IV. The $1\,\sigma$ constraints on parameters from the simulated mock data with the fiducial value $\Omega_k=-0.013$. \begin{center \begin{tabular}{c|c|c|c|c} \hline\hline models &$w_0$&$w_a$&$\Omega_K$&$\Delta\chi^2$\\ \hline fiducial model&$-1$&$0$&$-0.013$&$-$\\ \hline $CPL\,\oplus\, \Omega_k$&$-1.000^{+0.0270}_{-0.0280}$&$0.00\pm0.130$&$-0.0130^{+0.0017}_{-0.0020}$&$0$ \\ \hline $CPL$&$-1.068^{+0.0210}_{-0.0220}$&$0.481^{+0.074}_{-0.072}$& set to 0&$23$ \\ \hline Constant $w\,\oplus\, \Omega_k$&$-1.0000\pm0.0085$&set to 0&$-0.0127^{+0.0011}_{-0.0010}$&$0$ \\ \hline Constant $w$&$-0.9407^{+0.0057}_{-0.0058}$&set to 0& set to 0&$41$ \\ \hline\hline \end{tabular} \end{center} \end{table} We use the fiducial mock data to study the degeneracy between the EoS of dark energy and $\Omega_k$. When we simulate the future data, we choose the fiducial value of curvature $\Omega_k=-0.013$ which is allowed by current observational data. We summarize the numerical results in table IV. Firstly, we use the correct input cosmological model ``$CPL\,\oplus\,\Omega_k$'' to constrain parameters. By combining the future mock data from PLANCK, Euclid and supernovae, we obtain the constraints on the EoS parameters: $w_0=-1.000^{+0.0270}_{-0.0280}$ and $w_a=0.00\pm0.130$ at the 68\% confidence level. When comparing the constraints from the current observations, the future data constrain the EoS parameters much more stringently. The error bars of $w_0$ and $w_a$ are shrunk by a factor of 6. In figure \ref{fc_w0wa_omkd13} we plot the 2-dimensional constraint on the panel ($w_0$,$w_a$) (black solid lines). Meanwhile, the mock data also give very tight constraint on the curvature: $\Omega_k = -0.0130^{+0.0017}_{-0.0020}$ (68\% C.L.). \begin{figure}[t] \begin{center} \includegraphics[scale=0.6]{cpl_omkd13.eps} \caption{1 and 2-dimensional constraints on the panel ($w_0$,$w_a$) from the simulated mock data with the fiducial value $\Omega_k=-0.013$. The black solid lines are obtained by using the correct model which allows $\Omega_{k}$ varying. For comparison, we also use a biased model which fixes the curvature $\Omega_k\equiv0$ to constrain the EoS of dark energy, shown as the red dashed lines.\label{fc_w0wa_omkd13}} \end{center} \end{figure} Besides the correct input model, we also use a biased model to fit the mock data, in order to study the degeneracy between $\Omega_k$ and $w$. Here, we assume the flat universe ($\Omega_k\equiv0$). Due to the strong degeneracy between them, in this case the constraints of $w_0$ and $w_a$ are quite different from those of the correct input model. The $1\,\sigma$ constraints are $w_0=-1.068^{+0.0210}_{-0.0220}$ and $w_a=0.481^{+0.074}_{-0.072}$. One can see that the error bars are smaller than above, due to less free parameters in this case. More importantly, the mean values of $w_0$ and $w_a$ are shifted obviously, as shown in the red dashed lines of figure \ref{fc_w0wa_omkd13}. When considering the error bars, these mean values will be ruled out at more than $3\,\sigma$ confidence level, which implies that the future observations could break this degeneracy very well. We also use the constant $w$ model ($w_a\equiv0$) to fit the mock data again and plot the results in figure \ref{fc_womk_omkd13}. When varying $\Omega_k$, we obtain the constraints: $w=-1.0000\pm0.0085$ and $\Omega_k=-0.0127^{+0.0011}_{-0.0010}$ (68\% C.L.), which recover the fiducial model very well. However, when assuming the flat universe, due to the degeneracy, the $1\,\sigma$ constraint on $w$ is shifted: $w=-0.9407^{+0.0057}_{-0.0058}$, which will be excluded by the future data at more than $6\,\sigma$ confidence level. In this case, the minimal $\chi^2$ are very large, which implies that this kind of model does not fit the mock data well. \begin{figure}[t] \begin{center} \includegraphics[width=120mm,height=60mm]{cw_omkd13.eps} \caption{1-dimensional constraint on $w$ and 2-dimensional constraint on the panel ($\Omega_k$,$w$) from the simulated mock data with the fiducial value $\Omega_k=-0.013$. The black solid lines are obtained by using the correct model which allows $\Omega_{k}$ varying, while the red star denotes the best fit model when assuming the flat universe.\label{fc_womk_omkd13}} \end{center} \end{figure} \subsubsection{$\sum m_{\nu}$ and $w$} \begin{table TABLE V. The $1\,\sigma$ constraints on parameters from the simulated mock data with the fiducial value $\sum m_{\nu}=0.4\,$eV. \begin{center \begin{tabular}{c|c|c|c|c} \hline\hline models &$w_0$&$w_a$&$\sum m_{\nu}$&$\Delta\chi^2$\\ \hline fiducial model&$-1$&$0$&$0.4\,$eV&$-$\\ \hline $CPL\,\oplus\, m_{\nu}$&$-1.000\pm 0.0240$&$0.00\pm 0.110$&$0.400^{+0.020}_{-0.017}$&$0$ \\ \hline $CPL$&$-1.031\pm0.0140$&$0.589\pm 0.043$&set to 0&$40$ \\ \hline Constant $w\,\oplus\, m_{\nu}$&$-1.001^{+0.0130}_{-0.0120}$& set to 0&$0.400^{+0.011}_{-0.015}$&$0$ \\ \hline Constant $w$&$-0.8503^{+0.0037}_{-0.0038}$&set to 0& set to 0&$104$ \\ \hline\hline \end{tabular} \end{center} \end{table} \begin{figure}[t] \begin{center} \includegraphics[scale=0.6]{cpl_nud45.eps} \caption{1 and 2-dimensional constraints on the panel ($w_0$,$w_a$) from the simulated mock data with the fiducial value $\sum m_{\nu}=0.4\,$eV. The black solid lines are obtained by using the correct model which allows $\sum m_{\nu}$ varying. For comparison, we also use a biased model with the massless neutrino to constrain the EoS of dark energy, shown as the red dashed lines.\label{fs_w0wa_nud45}} \end{center} \end{figure} \begin{figure}[t] \begin{center} \includegraphics[width=120mm,height=60mm]{cw_nud45.eps} \caption{1-dimensional constraint on $w$ and 2-dimensional constraint on the panel ($w$,$\sum m_{\nu}$) from the simulated mock data with the fiducial value $\sum m_{\nu}=0.4\,$eV. The black solid lines are obtained by using the correct model which allows $\sum m_{\nu}$ varying, while the red star denotes the best fit model when assuming the massless neutrino model.\label{fs_cw}} \end{center} \end{figure} Similarly, we use the same method to study the degeneracy between $w$ and $\sum m_\nu$. When generating the simulated mock data, we choose the fiducial value of total neutrino mass $\sum m_{\nu}=0.4\,$eV, which is in agreement with the current constraint at $2\,\sigma$ confidence level. In figure \ref{fs_w0wa_nud45}, we plot the 2-dimensional constraints on $w_0$ and $w_a$ obtained by using different input cosmological models. Firstly, we consider the correct model (``$CPL\,\oplus\, m_{\nu}$'') and obtain very tight constraints on the EoS of dark energy and the total neutrino mass: $w_0=-1.000\pm0.0240$, $w_a=0.00\pm0.110$ and $\sum m_\nu=0.400^{+0.020}_{-0.017}$eV at the 68\% confidence level. The future measurements could give conclusive conclusion about the total neutrino mass. However, if assuming the massless neutrino ($\sum m_\nu\equiv0$), we will obtain the biased constraints on the EoS of dark energy: $w_0=-1.031\pm0.0140$ and $w_a=0.589\pm0.043$ (68\% C.L.), due to the strong degeneracy between $w$ and $\sum m_\nu$. The mean values are significantly different from the input fiducial model, especially for the parameter $w_a$. Consequently, the minimal $\chi^2$ of this case will be very large, which implies that this best fit model has been ruled out by the data at more than $3\,\sigma$ confidence level. We list the numerical results in table V. Again, we also use the constant EoS of dark energy to study this degeneracy. In the framework of the correct $w\,\oplus\,m_\nu$ model, the $1\,\sigma$ constraints are $w=-1.001^{+0.0130}_{-0.0120}$ and $\sum m_\nu=0.400^{+0.011}_{-0.015}$. When forcing $\sum m_\nu\equiv0$, the best fit model of $w$ is shifted to $w=-0.8503^{+0.0037}_{-0.0038}$, which is more than $10\,\sigma$ away from the fiducial value of $w$, as shown in figure \ref{fs_cw}. The future observations could break this degeneracy efficiently. \subsubsection{$n_s$, $\alpha$ and $r$} \begin{table TABLE VI. $1\,\sigma$ constraints on Inflationary parameters from the simulated mock data. \begin{center \begin{tabular}{c|c|c|c} \hline\hline models &$n_s$&$\alpha_s$&$\Delta\chi^2$ \\ \hline fiducial model&$1$&$-0.05$&$-$ \\ \hline correct~model&$1.0000\pm0.0010$&$-0.0502^{+0.0039}_{-0.0040}$&$0$ \\ \hline biased~model&$1.0052^{+0.0015}_{-0.0010}$&set~to~0&$78$ \\ \hline\hline models &$n_s$&$r$&$\Delta\chi^2$ \\ \hline fiducial model&$1$&$0.2$&$-$ \\ \hline correct~model&$1.0000^{+0.0011}_{-0.0010}$&$0.200^{+0.015}_{-0.020}$&$0$ \\ \hline biased~model&$0.9975{\pm 0.0009}$&set~to~0&$60$ \\ \hline\hline \end{tabular} \end{center} \end{table} In order to study the degeneracies between $n_s$ and $\alpha_s$, $r$, we simulate two mock data sets: one with $\alpha_s=-0.05$ and the other one with $r=0.2$, which are in agreement with the current data. In table VI we list the constraints of these Inflationary parameters from two mock data. Using the correct input model, we obtain very stringent constraints on $n_s$ and $\alpha_s$: $n_s=1.0000{\pm 0.0010}$, $\alpha_s =-0.0502^{+0.0039}_{-0.0040}$ (68\% C.L.), whose error bars are almost shrunk by a factor of 10. When assuming the spectral index is scale-independent ($\alpha_s\equiv0$), the constraint of $n_s$ changes: $n_s=1.0052^{+0.0015}_{-0.0010}$ at the 68\% confidence level. The mean value of $n_s$ is shifted significantly, when comparing with the obtained error bar. We compare these two results in figure \ref{fc_asns}. The obtained correlation coefficient between $n_s$ and $\alpha_s$ is $0.1$, which means this degeneracy does not show in the future data significantly. The similar situation is also found by the study about the degeneracy between $n_s$ and $r$. In figure \ref{fc_rns} we show the constraints on $n_s$ and $r$ from the correct and biased input models, respectively. When including the tensor perturbation mode, the $1\,\sigma$ constraints are: $n_s=1.0000^{+0.0011}_{-0.0010}$ and $r=0.200^{+0.015}_{-0.020}$. Neglecting the tensor perturbation, the constraint on $n_s$ is shifted to a lower value: $n_s=0.9975{\pm 0.0009}$. We obtain the correlation coefficient between $n_s$ and $r$ is $-0.05$, which implies that the degeneracies between $n_s$ and $r$ is not notable for the future data sets. \begin{figure}[t] \begin{center} \includegraphics[width=120mm,height=60mm]{lcdm_asns.eps} \caption{1-dimensional constraint on $n_s$ and 2-dimensional constraint on the panel ($n_s$,$\alpha_s$) from the simulated mock data with the fiducial value $\alpha_s=-0.05$. The black solid lines are obtained by using the correct model which allows $\alpha_s$ varying, while the red star denotes the best fit model when fixing $\alpha_s=0$.\label{fc_asns}} \end{center} \end{figure} \begin{figure}[t] \begin{center} \includegraphics[width=120mm,height=60mm]{lcdm_rns.eps} \caption{1-dimensional constraint on $n_s$ and 2-dimensional constraint on the panel ($n_s$,$r$) from the simulated mock data with the fiducial value $r=0.2$. The black solid lines are obtained by using the correct model which allows $r$ varying, while the red star denotes the best fit model when neglecting the tensor perturbation mode.\label{fc_rns}} \end{center} \end{figure} \subsection{Conclusion and discussion} In this paper we perform global fitting analyses on serval cosmological models and present the constraints on cosmological parameters from the latest astronomical observations, including WMAP7, BAO measurement from the SDSS-II survey and supernovae ``Union2.1'' sample. We find that, with different input theoretical models, the constraints on some cosmological parameters can be very different, such as the equation of state of dark energy, the total neutrino mass and the index of primordial power spectrum, due to the strong degeneracies among these parameters. With the high quality mock data, such as future SN, CMB and LSS observations, we further study the impact of these degeneracies on the constraints of cosmological parameters. Using these accurate future data, the obtained results show that, some degeneracies can not be neglected, such as the EoS of dark energy and the curvature, the EoS and the massive neutrino. Ignoring them forcibly will lead to seriously biased constraints. \section*{Acknowledgements} We thank Xinmin Zhang and Gong-Bo Zhao for helpful discussion. We acknowledge the use of the Legacy Archive for Microwave Background Data Analysis (LAMBDA). Support for LAMBDA is provided by the NASA Office of Space Science. The calculation is taken on Deepcomp7000 of Supercomputing Center, Computer Network Information Center of Chinese Academy of Sciences. This work is supported in part by the National Science Foundation of China under Grant Nos. 11033005, by the 973 program under Grant No. 2010CB83300, by the Chinese Academy of Science under Grant No. KJCX2-EW-W01.
\section{Self-energy scaling} In a Fermi liquid, $\mathrm{Im}\Sigma(\omega,T) \propto \left[\omega^2+(\pi T)^2\right]$ and therefore the self-energy obeys the following scaling in $\omega/T$: \begin{equation} -\frac{D\mathrm{Im} \Sigma}{T^2} = A \left[\pi^2+ (\omega/T)^2\right]. \label{eq:scaling} \end{equation} In Fig.~\ref{fig:FLscale}, we show $-\mathrm{Im}\Sigma(\omega,T)/T^2$ as a function of $\omega/T$ for different temperatures. As the temperature is lowered below $T_\mathrm{FL}\approx0.01D$, the curves collapse on a parabola, confirming the expected Fermi-liquid scaling law. Finding such a scaling is actually a stringent test on the numerical data. Very precise quantum Monte Carlo data (analytically continued with Pad\'e approximants) and a strict control of the chemical potential were needed to obtain these results. \begin{figure} \includegraphics[width=1\columnwidth]{fig_s1.pdf} \caption{The self energy scaling as $\omega/T$ for the doped Hubbard model considered in the main text (Bethe lattice, $U=4D$, doping $\delta=0.2$). At low temperatures self energy follows Eq.~\ref{eq:scaling}. $A$ is found to be close to $Z^{-2}\sim \delta^{-2}$. (For doping 0.2, $Z=0.22$.) \label{fig:FLscale}} \end{figure} As the temperature is raised, the positive-frequency side quickly deviates from the scaling form, revealing a particle-hole asymmetry already for temperatures above $T/D=0.005$. The deviations appear at a scale $\omega_+ \approx \pi T_{\mathrm{FL}}$. On the negative-frequency side the self-energy follows the quadratic behavior much more robustly and deviates from the scaling function only at about $T/D = 0.02$. Notice that the corrections on the $\omega>0$ side grow linearly with temperature. This leads to a Seebeck coefficient which is linear in $T$ at low temperatures, but with an enhanced slope as compared to the result one would get if these corrections were neglected. The transport probes an energy window of a few (say from -5 to 5) $k_B T$ . In this energy window the self-energy starts to deviate appreciably at $T/D = 0.01$. This is where the resistivity (within the precision of our data) visually departs from the $T^2$ law. \section{Temperature evolution of momentum-resolved spectra} In Fig.~\ref{fig:contour} we plot the temperature evolution of the momentum-resolved spectra using a color-map where bright (dark) colors indicate high (low) values of $A_k(\omega)$. At low temperatures, the data display two peaks corresponding to: the lower Hubbard band (LHB) which disperses around $\omega_{\mathrm{LHB}}\sim -\mu_0$ and the quasiparticle peak (QP) in the vicinity of $\omega \sim 0$. $\mu_0$ is the effective chemical potential at $T=0$. The upper Hubbard band (UHB) centered at the energy $\omega_{\mathrm{UHB}}\sim U-\mu_0$ lies above the energy range displayed in the plot. The lowest temperature data $T/D=0.0025$ show a very sharp QP peak around $\omega=0$, which is rapidly broadened as the frequency is increased. At very small frequencies, the slope of the dispersion is found to be approximately 5 times smaller than the bare one, as dictated by $Z\approx 0.2$. On the negative-frequency side this holds almost until the bottom of the band. On the positive-frequency side, instead, the kink at $\omega_+$ is rapidly encountered. Above this kink, the slope of the band dispersion increases to about half the bare dispersion slope. The LHB, seen clearly for occupied states $k<k_F$, disperses at a slope close to that of the bare dispersion. \begin{figure*} \includegraphics[width=2\columnwidth]{Fig_Spec_ContourMap_hq.jpg} \caption{Contour map of momentum-resolved spectra $A(\epsilon_k,\omega)$ at various temperatures for the same parameters as the data in the main text: doping $\delta=0.2$ and $U/D=4.0$. \label{fig:contour}} \end{figure*} As the temperature is increased, the QP band broadens and becomes more dispersing. It becomes therefore progressively more difficult to resolve it from the LHB band. Nevertheless, for temperatures well above $T_\mathrm{FL}$ and $T_*$ (four leftmost panels) one can still clearly distinguish the QP band from the LHB. The maximum of the spectra is also indicated (lines). This maximum has a discontinuity at a point where the maximal value in the QP band becomes larger than the maximal value reached in the LHB band. Dashed and solid lines are used to denote the maximum in the QP and LHB band, respectively. Above $T/D=0.2$ the maximal value does not have a discontinuity anymore and the signature of the quasiparticles is only visible as a kink in the dispersion. This marks the onset of the bad-metal regime. Note that $T/D=0.2=\delta$ corresponds to the Brinkman-Rice scale. At the highest temperature $T/D=1.0$ the kink is not seen anymore. The QP band crosses the Fermi energy at different momenta as the temperature increases. Identifying the Fermi surface with the momenta at which the spectral intensity of the QP band is maximal leads to the conclusion that the Fermi volume inflates as the temperature is increased. Note that the number of particles is fixed so that with this identification of the Fermi surface the Luttinger theorem is only obeyed at very low temperatures. To elaborate on this, we plot on Fig.~\ref{fig:Ak0} the momentum-distribution curve at $\omega=0$, \begin{equation} A(\epsilon_k,0) = \frac{1}{\pi}\frac{-\mathrm{Im}\Sigma(0,T)}{(\mu-\mathrm{Re}\Sigma(0,T)-\epsilon_k)^2+(\mathrm{Im}\Sigma(0,T))^2} \label{eq:mdc} \end{equation} for several temperatures. This spectral function has the shape of a Lorentzian centered at $\mu-\mathrm{Re}\Sigma(0,T)-\epsilon_k$ with a width $\mathrm{Im}\Sigma(0,T)$. \begin{figure} \includegraphics[width=1\columnwidth]{fig_s3.pdf} \caption{The evolution of momentum-distribution curves at the Fermi level Eq.~(\ref{eq:mdc}). \label{fig:Ak0} } \end{figure} At very low temperatures a sharp peak lies at the chemical potential, fulfilling the Luttinger theorem. When the temperature increases, the peak moves to higher momenta. An alternative way to track this change is to look at the renormalized chemical potential $\mu_\mathrm{eff} = \mu-\mathrm{Re}\Sigma(0,T)$ as a function of the temperature as shown in Fig.~\ref{fig:mu_eff}. In the Fermi-liquid regime, $\mu_\mathrm{eff}$ essentially follows the noninteracting chemical potential $\mu_0$ (shown with a dashed line). At higher temperatures $\mu_\mathrm{eff}$ rapidly increases. \begin{figure} \includegraphics[width=1\columnwidth]{fig_s4.pdf} \caption{Temperature dependence of $\mu_\mathrm{eff}=\mu(T)-\mathrm{Re}\Sigma(\omega=0,T)$. The noninteracting chemical potential $\mu_0$ is shown by a dashed line. \label{fig:mu_eff}} \end{figure} An important lesson here is that observing a well-defined QP peak does not imply that the system has reached the Fermi liquid regime. As Fig.~\ref{fig:mu_eff} shows, the chemical potential in this intermediate-temperature metal can be quite far from the Fermi energy, despite a signature of well-distinguishable resilient QPs. \section{Thermopower at high temperatures and comparison to approximate formulas} In Fig.~\ref{fig:seebeck_highT} we show the Seebeck coefficient over a larger temperature window. The Seebeck coefficient calculated using the Kubo formula is plotted with a thick line and compared to various estimates. The Kelvin formula $\partial \mu/\partial T$ (using $\mu$ calculated within DMFT) overestimates the magnitude of the thermopower in the low-$T$ regime but is a good approximation of the exact result above $T_*\approx 0.08D$. For comparison, we also plot the corresponding atomic estimate using the Kelvin formula, but using the $\mu$ obtained in the atomic limit. We note that the two expressions essentially match above $T/D=1$. Note that the Kubo result starts to deviate significantly from the atomic estimate only when entering the resilient QP regime. Finally, we plot Heikes estimates. The results obtained from the DMFT chemical potential using NRG and continuous-time interaction expansion Monte Carlo (CTINT) are plotted (full green line and symbols). Heikes formula is found to approximate the thermopower worse than the Kelvin formula. For comparison, also the atomic Heikes estimates (thick dashed) as well as the asymptotic $U\to0$ and $U\to \infty$ Heikes values (horizontal lines) are shown. \begin{figure} \includegraphics[width=1\columnwidth]{fig_s5.pdf} \caption{Seebeck coefficient calculated using the exact Kubo formula (thick black line) compared to the approximate Kelvin formula (red line) and Heikes formula (green line and green symbols). The atomic Kelvin estimate (dotted) and Heikes estimate (thick dashed) interpolate between the asymptotic $U\to\infty$ and $U\to 0$ Heikes values.\label{fig:seebeck_highT}} \end{figure} \end{document}
\section{Introduction} High energy heavy ion collisions provide a unique opportunity to investigate the properties of nuclear matter at high energy density, a few GeV/fm$^{3}$, which is several times the energy density of a proton. At these temperatures, $\gtrsim 170$ MeV, matter undergoes a transition between the hadronic phase and the quark-gluon plasma phase (QGP) \cite{Van:2002tm,McLarren:1986qg}. In the prevailing view, the strongly interacting matter validates the use of thermodynamics \cite{Stachel:1989th,Braun:1995th,Braun:1996th,Andronic:2011yq,Rafelski:1991th,Davidson:1991th,Sollfrank,Wheaton:2004qb} to describe the chemistry and hydrodynamics \cite{Kolb:2003dz} to describe the evolution. A typical assumption is that the hadronic chemistry is described by chemical freeze-out at a single temperature. This temperature is also associated with the QGP-hadron transition, or hadronization. Although hadrons certainly collide a few more times after hadronization, many models typically ignore the effects of hadronic-stage interactions on the chemistry. Particle ratios are often fit by a single temperature and a few chemical potentials to account for non-zero conserved charges such as strangeness, baryon number and electric charge. Additionally, fugacities have been added to account for the inability of the system to produce the requisite number of up, down and strange quarks for equilibrium \cite{Torrieri:2005va}. These pictures are usually based on the idea that, due to the strong interactions between quarks, all partitions into various hadron species are thermally sampled during hadronization, and that the evolution of chemical abundances after hadronization comes only from decays. However, the inferred temperature could be misleading if chemical reactions during the hadronization stage significantly change particle ratios. Of course, the chemical make-up does not completely freeze-out --- hadrons indeed undergo chemical reactions during the hadronic stage. The effects of such evolution have been studied in two contexts. First, hadronic cascades (microscopic simulations that follow the trajectories of individual hadrons) such as URQMD \cite{URQMD} include inelastic processes. Although these models include annihilation processes, e.g., $p\bar{p}\rightarrow n\pi$ \cite{Karpenko:2012yf,Steinheimer:2012bn,Steinheimer:2012rd,Becattini:2012sq,Becattini:2012xb}, few microscopic models have included the regeneration processes and therefore violate detailed balance. Inverse processes have been added to the hadronic HSD model, where $2\leftrightarrow 3$ processes such as $\rho\rho\pi\leftrightarrow p\bar{p}$ are included self consistently \cite{Cassing:2001ds}. Concomittantly, $2\leftrightarrow 3$ processes have also been incorporated into partonic simulations \cite{Xu:2004mz}. The particular process, $\bar{p}p\rightarrow \rho\rho\pi$, allows an annihilation to ultimately produce 5 mesons after the $\rho$ mesons decay. These incorporations into the HSD model have been used to investigate the possibility of creating the observed number of antibaryons through bottom-up processes, mostly at SPS, AGS and SIS energies. At higher energies, if one assumes that a chemically equilibrated number of baryons and antibaryons are created at hadronization, the issue becomes one of whether the regeneration process significantly dampens the annihilation rate during the hadronic stage. Implementing regeneration into cascade codes , as was done in HSD, requires significant investment, both in development and in the numerical cost of running more complicated codes. A second class of approaches involves solving for the chemical rates of a kinetically equilibrated gas. Given that there are hundreds of resonances, such models typically assign fugacities to a few numbers such as effective pion number, or the net number of baryons plus antibaryons, and then solve for the evolution of the fugacities \cite{Scott:1999ha,Chung:1997ch,Rapp:2002ha,Rapp:2000gy,Huovinen:2003sa}. In \cite{Rapp:2000gy} it was shown that regeneration rates, e.g. $n\pi\rightarrow p\bar{p}$, were important. Related issues were pursued by considering regeneration through Hagedorn states in \cite{NoronhaHostler:2007jf,Greiner:2004vm,NoronhaHostler:2009cf}, where it was estimated that chemical equilibration times were only fast enough to maintain equilibrium for temperatures $\gtrsim$ 170 MeV. Chemical evolution was superimposed onto a hydrodynamic calculation in \cite{Huovinen:2003sa}, where regeneration processes were included in a matter consistent with reproducing chemical equilibrium. In their results, annihilation and regeneration reduced baryon yields by 15-20\%. In all the calculations mentioned above, it is assumed that chemical equilibrium is maintained until some point, and that this temperature is sufficiently low that the system can be approximated as a hadron gas. This temperature, which we will label $T_0$, is sometimes associated with the hadronization temperature, though that is not necessary. Inferred temperatures near 170 MeV are common, even though the density of a hadron gas at such temperatures would suggest densities of one hadron per fm$^3$, which is roughly the inverse volume of a hadron. Further, lattice calculations show that the system is undergoing significant changes already by that time. The scalar quark condensate and the Polyakov loop are already significantly changed relative to vacuum values for temperatures $\sim 160$ MeV \cite{Bazavov:2013yv}. Fluctuations of conserved charges measured on the lattice show that hadron-like states exist up to somewhat higher temperatures, $\sim 200$ MeV \cite{Ratti:2011au}{Bazavov:2013dta}. We proceed here with calculations assuming $T_0=170$ MeV, even though hadrons at this temperature may have a rather different character than those in the vacuum. The shape of pion, kaon and proton spectra can be well described by hybrid hydrodynamic/cascade models that assume chemical equilibrium at the hadronization temperature, and ignore any subsequent chemical evolution beyond resonance decays and reformation \cite{Song:2011hk,Novak:2013bqa}. However, such model's proton/pion ratio tends to significantly differ from experimental measurements from the PHENIX Collaboration at RHIC and from the ALICE Collaboration at the LHC. PHENIX ratios \cite{Adler:2003cb} were $\sim 35$\% below hybrid model predictions that fit the yields of pions and kaons in \cite{Song:2011hk,Novak:2013bqa} where baryon annihilations had been neglected. More consistent shortcomings are seen in Pb+Pb collisions at the LHC measured by ALICE \cite{Abelev:2012wca}. The $p/\pi$ ratio is nearly half of what one would expect from freezeout at hadronization temperatures \cite{Steinheimer:2012bn}, and are $\sim 50\%$ lower than the hydrodynamic results in \cite{Paatelainen:2012at}. Using a microscopic model, a reduction of approximately this size was seen \cite{Karpenko:2012yf} and in the erratum to \cite{Song:2011hk}, but those calculations ignore the inverse rates, which according to \cite{Rapp:2000gy} largely cancel the annihilation rates. In \cite{Steinheimer:2012rd}, the inverse process was crudely estimated at the level of 8\%, with this value then being used to represent the error associated with the lack of detailed balance in their model. The principal goal of this paper is to solve for the chemical rates of a wide range of hadronic species where baryon annihilation and regeneration are treated in a manner consistent with detailed balance, and to understand the relative importance of annihilation and regeneration. This is similar in spirit to what was outlined in \cite{Rapp:2000gy} and performed in \cite{Huovinen:2003sa}, but in more detail and with a larger number of resonances considered. The chemical evolution is also followed until interactions fully cease. We will determine whether in a consistent calculation annihilations might account for a $\sim 40-50$\% reduction in the final-state $p/\pi$ ratio measured in heavy-ion collisions at high energy relative to the final-state ration calculated without annihilation processes. Further, we discuss the importance of accounting for baryon annihilations when estimating the chemical freeze-out temperature from final-state yields. As a secondary goal, we wish to see whether strange baryons are affected at the same level. In order to concentrate on chemistry, we employ a simplified model of the space-time evolution. We consider the evolution of a single hydrodynamic cell, where the volume changes as a function of time according to a simple prescription. The parameterization is based on the one-dimensional Bjorken expansion \cite{Bjorken:1983}, but is modified to account for transverse expansion. The parameterization is chosen to roughly match what happens in a three-dimensional hydrodynamic calculation of a central Pb+Pb collision at the LHC. In a more realistic model, one would consider the whole ensemble of hydrodynamic cells, each of which would evolve differently. Various species lose local kinetic equilibrium with one another when temperatures approach 140 MeV, as heavier species like protons cool faster than pions \cite{Pratt:1998gt}. Additionally, pions would begin to flow with different velocities, and leave the collision region while the embers would become relatively more baryon rich \cite{Sorge:1995pw}. Accounting for these non-equilibrium aspects requires microscopic simulations. Nonetheless, this picture should be adequate to gauge the significance of annihilation and regeneration. \section{Description of Calculation} We model the hadronic stage by first assuming that all resonances are initially populated according to chemical equilibrium with a temperature $T_0$, set by default to 170 MeV, which is in the neighborhood of where hadronization takes place. All chemical potentials are set to zero. The list of 319 resonances taken from the Particle Data Book \cite{particledatabook} extends to masses up to 2.25 GeV. Although one should have a non-zero baryonic chemical potential, even at high RHIC and LHC energies, it will be ignored for this study given that the our goal is to gauge the importance of annihilation effects. After the initial time, we assume that collisions are sufficiently frequent to preserve local kinetic equilibrium, even though inelastic collisions are insufficient for maintaining chemical equilibrium. Each species is assigned a chemical potential, $\mu_i$. Each $\mu_i$ is initially set to zero, and if chemical equilibrium were maintained each $\mu_i$ would remain zero. Ignoring Bose effects, which are a $\sim 10\%$ correction for pions, thermal results within the context of the grand canonical ensemble for number densities, pressure, energy density and entropy are: \begin{eqnarray}\label{differentiation} n_{i}&=&\frac{g_i}{2\pi^2} \left(m_i^2TK_0(m_i/T)+2m_iT^2K_1(m_i/T)\right)e^{\mu_i/T}, \nonumber \\ P&=&\sum_i n_iT,\\ \nonumber \epsilon &=&\sum_i\frac{g_i}{2\pi^2} \left[3m_i^2T^2K_0(m_i/T)+\left(m_i^3T+6m_iT^3\right)K_1(m_i/T)\right]e^{\mu_i/T},\\ \nonumber s&=&\frac{P+\epsilon-\sum_i\mu_in_i}{T}, \end{eqnarray} where $g_{i}$ and $\mu_{i}$ are the degeneracy and chemical potential of the hadron species $i$, and $E_{i}=\sqrt{p^{2}+m_{i}^{2}}$ where $m_{i}$ is the particle mass and $K_{n}$ is the Bessel function of order $n$. For a chemically equilibrated system, the chemical potentials, $\mu_i$, are zero. For a system of massless particles in an isentropic expansion, the yields can be frozen and $\mu_i$ will remain zero. However, for massive particles the chemical potentials become positive if yields are frozen. This can be understood by considering the fact that the entropy per particle with zero chemical potential is near 3.5 for massless particles, but increases as a function of $m/T$. Even pions, the lightest mesons, develop significantly positive chemical potentials if the chemical evolution is frozen or if the evolution is confined to decays \cite{Greiner:1993jn,Scott:1999ha}. As will be seen later in this paper, chemical potentials remain positive even after decays, e.g. $\Delta\rightarrow N\pi$, and annihilations, e.g. $p\bar{p}\rightarrow 5\pi$ are included. Baryon annihilation reduces baryon chemical potentials, but again are insufficiently fast to maintain equilibrium. To return the pion chemical potential to zero, pion absorbing processes, e.g. $pp\pi_0\rightarrow p\Delta^+\rightarrow pp$, are required. Due to conservation of g-parity, there are only a few reactions in the meson sector that can reduce the final number of pions relative to the number one would get from decays alone. For example, $\rho+\pi\rightarrow\pi\pi$ violates g-parity. Reactions such as $\rho\rho\rightarrow\pi\pi$ or $\omega\pi\rightarrow\pi\pi$ conserve g-parity but are fairly rare. The baryon sector is a better candidate for affecting the final pion number since g-parity is not a constraint for reactions involving baryons. For instance $\Delta N\rightarrow NN$ lowers the final pion number by one. Such reactions might well reduce the final-state number of pions at the 5-10\% level. As will be shown here, pions become overpopulated by a factor $e^{\mu_\pi/T}\approx 1.75$ when the temperature has cooled to approximately 100 MeV. Thus, pion-absorbing reactions are far from able to maintain equilibrium, but they might also be non-negligible since chemical yields are now being evaluated at the 10\% level. Baryons comprise only $\approx 10\%$ of final-state hadrons at the LHC or at the highest RHIC energies, but represent a much higher fraction of the hadrons at SPS energies. Although such pion-absorbing processes are not considered here, they deserve further study, especially for analyses of lower-energy collisions. Since we are interested in simply gauging the effects of annihilation, we consider a simplified picture of the expansion. In a one-dimensional boost-invariant expansion \cite{Bjorken:1983}, the volume of a fluid element would increase proportional to the time $\tau$. However, such a picture neglects transverse expansion. To crudely account for transverse expansion, we assume that the volume of the element increases as \begin{equation} \label{eq:OmegaOfTau} \Omega(\tau)=\Omega(\tau_0)\frac{\tau}{\tau_0}\frac{\tau_\perp^2+\tau^2}{\tau_\perp^2+\tau_0^2}, \end{equation} for times $\tau$ greater than the hadronization time $\tau_0$. By viewing the evolution of the density and temperature of hydrodynamic expansions from \cite{Vredevoogd:2012ui}, the parameters were chosen to be: $\tau_0=8$ fm/$c$, $\tau_\perp=6.5$ fm/$c$. This parameterization produced density vs. time curves that were similar to those seen in full hydrodynamic models. Based on the assumption above, the continuity equations can be easily be solved in Bjorken coordinates should all chemical rates be turned off, \begin{eqnarray} n_{i}(\tau)=n_{i}(\tau_{0})\frac{\Omega(\tau_0)}{\Omega(\tau)}, \nonumber \\ s(\tau)= s(\tau_{0})\frac{\Omega)(\tau_0)}{\Omega(\tau)}. \end{eqnarray} The evolution equations for the chemistry are modified in the presence of chemical rates, \begin{equation} \label{eq:dndtau} \frac{d}{d\tau}\left[\Omega(\tau) n_i(\tau)\right]=\Omega(\tau) R_i(\tau), \end{equation} where $R_i$ is net production/annihilation rate per unit $d^4x$ of particles of species $i$. The net rates, $R_i$, are all zero at equilibrium, but as the system cools the system loses equilibrium and the rates become non-zero (usually negative) to push toward equilibrium. Once the system loses chemical equilibrium, entropy can be generated, \begin{equation} \label{eq:dsdtau} \frac{d}{d\tau}\left[\Omega(\tau) s(\tau)\right]=-\Omega(\tau)\sum_i R_i(\tau)\frac{\mu_i(\tau)}{T(\tau)}. \end{equation} Given the entropy and number densities, one can solve for the temperature and chemical potentials. Thus, Given the rates $R_i$, Eq.s (\ref{eq:dndtau}-\ref{eq:dsdtau}) allow one to solve for the evolution of the number and entropy densities, or equivalently the chemical potentials and temperatures, as a function of $\tau$. A multidimensional Newton's method is used to numerically determine $\mu_i$ and $T$ from $n_i$ and $s$. One contribution to $R_i$ is resonances decays. For the reaction $A\to a_{1}+a_{2}+\cdots+a_{n}$, the contribution to the growth rate for resonance of type $A$ is \begin{equation} R_A(\tau)=-\left(1- \exp\left\{\frac{\mu_{a_{1}}+\mu_{a_{2}}+\cdots+\mu_{a_n}-\mu_{A}}{T}\right\}\right)\left\langle\Gamma_{A\rightarrow a_1\cdots a_n}\right\rangle n_{A}(\tau). \end{equation} Here $\langle\Gamma\rangle$ is the decay rate of the resonance, which is the nominal width divided by the average Lorentz time dilation factor, \begin{equation} \langle\Gamma\rangle=\frac{\Gamma}{(2\pi)^3n_A}\int(d^3p/E_A)~m_A f(p). \end{equation} Partial widths are taken from the particle data book \cite{particledatabook}. The contribution from the inverse reaction, $a_1+\cdots a_n\rightarrow A$, is accounted for by the factor $\exp[(-\mu_A+\mu_1\cdots +\mu_n)/T]$. This forces $R_A$ to go to zero when the chemical potentials balance. This channel also contributes to the growth rates for $a_1\cdots a_n$ with the same magnitude but opposite sign. All the hadronic decay channels from the particle data book are included in the calculation. The second type of chemical reaction to be considered is that from baryon annihilation, $A+\bar{B}\rightarrow a_1+\cdots a_n$. The contribution to the growth rate of $A$ in this case is \begin{equation} \label{eq:annihilationdamping} R_A(\tau)=-\left(1- e^{(\mu_{a_{1}}+\mu_{a_{2}}+\cdots+\mu_{a_{n}}-\mu_A-\mu_B)/T}\right) \left\langle\sigma_{A+\bar{B}\rightarrow a_1\cdots a_n}v_{\rm rel}\right\rangle n_{A}(\tau)n_B(\tau). \end{equation} The average $\langle \sigma v_{\rm rel}\rangle$ for a relativistic gas is \begin{equation} \left\langle\sigma v_{\rm rel}\right\rangle= \frac{1}{(2\pi)^6n_An_B}\int (d^3p_a/E_a)(d^3p_b/E_b)~f_a(p_a)f_b(p_b)\sigma(s)\sqrt{(p_a\cdot p_b)^2-m_a^2m_b^2}. \end{equation} We apply a simplified form for the $p\bar{p}$ annihilation cross section that is accurate for $P_{\rm lab}>100$ MeV/$c$ \cite{Wang:1998sh,ppbardata}, \begin{equation} \label{eq:xsection} \sigma=67~P_{\rm lab}^{-0.7} {\rm mb}. \end{equation} Without particularly good justification, we further assume all baryon-antibaryon pairs have the same annihilation cross section as $p\bar{p}$. Averaged over momentum, the average $\langle \sigma v_{\rm rel}\rangle$ is near 5 fm$^2\cdot c$ for $p\bar{p}$. Other inelastic processes, e.g., $\Delta+N\rightarrow NN$, are ignored for this study as we are focusing on baryons. Such processes can change $n_\pi$ at the level of 10\% \cite{Scott:1999ha,bao:1995for}. Experimental data show that the most likely number of pions in $p\bar{p}\rightarrow N\pi$ is $\sim 5$, and to simplify the calculation we assume that all annihilations proceed through the $5~\pi$ production channel. For pair annihilation with nonzero strangeness, we employ the minimal rule for kaon production, i.e, the number of kaons in the final state equals the initial strangeness. For example, if the initial strangeness is unity, then final production of annihilation will have one kaon. We still assume final production has five particles on average. The number of pions, $n_{\pi}$, will be $5-n_{K}$. Before displaying numerical results, we summarize the assumptions in our calculation. We consider a system with initial charge set to zero, and at chemical equilibrium so that $\mu_i=0$ for all species. The initial temperature is set to 170 MeV, and the initial time is set to $\tau_0=6$ fm/$c$. The volume of the cell scales as in Eq. (\ref{eq:OmegaOfTau}). Only decays through the strong interaction are considered since both electromagnetic and weak decays are negligible on these time scales. The calculation considered 319 resonances, with masses up to 2.25 GeV/$c^2$. If experimental data or competing models include feed-down from weak decays, the model used here can easily be modified for the purposes of a consistent comparison. \section{Results} Figure \ref{fig:alldens} shows the evolution of the densities of pions, kaons, protons, Lambdas, Sigma, Cascade and Omega baryons. Results are displayed both for when annihilations are enabled and disabled. When annihilations are disabled, the yields of stable particles increase due to feeding from the decays of unstable particles. Decays increase the number of protons by approximately 50\%, mainly through delta decays. The effect of baryon annihilation is to lower the baryon yields by $\sim$40\% relative to the yields without annihilation, to increase pion yields by $\sim$10\% and to increase the kaon yields by $\sim$5\%. Annihilations reduce the yields of strange baryons by approximately the same factor as protons. At $\tau_0$ annihilations play no role because the inverse process exactly cancels annihilation at equilibrium. However, within a few fm/$c$, the system is no longer at chemical equilibrium and the curves with and without annihilation diverge. \begin{figure}[htbp] \centerline{\includegraphics[width=0.5\textwidth]{pan_fig1}} \caption{\label{fig:alldens} (color online) Hadronic densitiies, scaled by the volume $\Omega(\tau)/\Omega(\tau_0)$, for the $\pi$ and $K$ mesons, and for the $p$, $\Lambda$, $\Sigma$, $\Xi$, and $\Omega$ baryons. In the absence of annihilation, the proton yield increases due to the decay of resonances like the $\Delta$. By adding annihilation, all baryon yields fall by $\sim 40$\%. Meson yields are modestly increased by the annihilation processes. If regeneration processes are ignored, the fraction of baryons that are annihilated increases to $\sim 50\%$ } \label{Figure1} \end{figure} \begin{figure}[htbp] \centerline{\includegraphics[width=0.5\textwidth]{pan_fig2}} \caption{\label{fig:muT} (color online) The annihilation rate (upper panel) as a function of time, beginning at the point when chemical equilibrium is lost, at $\tau_0=8$ fm/$c$. Regeneration blocks the annihilation for the first one or two fm/$c$, but after four or five fm/$c$ becomes negligible. When regeneration is included nearly half the annihilation occurs after the temperature has fallen below 125 MeV. The evolution of the temperature and chemical potentials for three species as a function of time in the middle two panels. The lower panel shows the fraction of the annihilation rate that is uncanceled by regeneration. In the latter stages, regeneration becomes negligible. } \end{figure} As the system cools, chemical equilibrium is lost and the chemical potentials become non-zero. Figure \ref{fig:muT} shows the evolution of the temperature and chemical potentials as a function of time. The chemical potentials develop more strongly for more massive particles since their yields are more sensitive to temperature. The lower panel demonstrates the degree to which the inverse processes cancel annihilation by considering the ratio $(1-e^{(-2\mu_p+5\mu_\pi)/T})$ which represents the fraction of annihilations canceled by regeneration in Eq(\ref{eq:annihilationdamping}). At the initial time, chemical populations are equilibrated and the inverse processes, e.g. $5\pi\rightarrow p\bar{p}$, exactly cancel the annihilation processes. However, the cancelation rapidly disappears, and by the end of the reaction, regeneration is negligible. From the upper panel one can see that annihilation continues for a significant time, and falls roughly proportional to the density, $\sim 1/\tau^3$. After accounting for regeneration, half the annihilation occurs after temperatures fall below 125 MeV. In \cite{Huovinen:2003sa} consistent chemical rates were used for a hydrodynamic calculation, but in that calculation freezeout was assumed for a temperature of 120 MeV. Such a choice, at least for LHC modeling, would miss a significant fraction of the annihilation. This may explain part of the reason why in \cite{Huovinen:2003sa} only 15-20\% of the baryons were annihilated, while in this calculation the rate was closer to 40\%. A common use of particle ratios is to determine the chemical freeze-out temperature. Although analyses of this type involve a global fit, the temperature is mainly determined by ratios between the heaviest and lightest particles since they are the most sensitive to the temperature. The $p/\pi$ ratio is especially important. Since such analyses, c.g. \cite{Andronic:2011yq}, typically ignore baryon annihilation it is instructive to see the degree to which annihilation processes distort the extraction of a chemical freezeout temperature. Figure \ref{fig:poverpi} displays the final $p/\pi^+$ ratio as a function of the initial temperature for calculations with and without baryon annihilation. For initial temperatures different than 170, the initial time $\tau_0$ was adjusted so that the temperature vs time trajectory would closely match that of the default calculation with $T_0=$170 MeV and $\tau_0=8$ fm/$c$. In Fig. \ref{fig:poverpi} the calculations with annihilation also include regeneration. The ALICE collaboration at the LHC measured $p/\pi^+=0.46\pm 10\%$ \cite{Abelev:2012wca}, and is illustrated with a grey band in Fig. \ref{fig:poverpi}. Whereas this ratio would correspond to a chemical freeze-out temperature of 145 MeV if baryon annihilation were ignored, if one accounts for the effects of baryon annihilation and regeneration, one becomes fairly insensitive to $T_0$, and any values greater than $\gtrsim 150$ MeV seem acceptable. \begin{figure} \centerline{\includegraphics[width=0.6\textwidth]{pan_fig3}} \caption{\label{fig:poverpi}(color online) The final $p/\pi^+$, $\Xi^-/\pi^-$ and $\Omega^-/\pi^-$ ratios as a function of $T_0$, the temperature at which chemical equilibrium is lost. Calculations with (solid red) and without baryon annihilation differ substantially. If one were to use the experimental $p/\pi^+$ ratio measured by ALICE to infer $T_0$, including annihilation would change the inferred value from $\sim 145$ MeV to a broad range of higher temperatures. If the annihilation cross sections were reduced for those baryons with a higher strangeness content, using the factor $\alpha=0.75$ in Eq. (\ref{eq:sreduction}), the resulting ratios are modestly higher for the multistrange baryons. } \end{figure} Figure \ref{fig:poverpi} also shows two hyperon to pion ratios, $\Xi/\pi$ and $\Omega/\pi$. Using the same cross section for the annihilation of any two baryon species, independent of the isospin, spin, or strangeness content of the annihilating baryons, the annihilation fraction turned out to be largely independent of strangeness. Of course, there is little reason to think that the annihilation cross section for $\Omega-\bar{\Xi}$ should be the same as for $p\bar{p}$. It is known that $\bar{p}n$ cross sections are similar to $\bar{p}p$ \cite{pbarn}, so one might feel justified in assuming that annihilation only weakly depends on isospin. However, virtually nothing is known about how it might be affected by strangeness. There are model-dependent calculations \cite{kovacs} or \cite{Kapusta:2002pg}, which are unconstrained by data. For example, in \cite{Kapusta:2002pg} cross sections are calculated in a simple field theory incorporating SU(3) flavor symmetry, and the chemical relaxation times for various hyperons vary on the order of a few tens of percent from that for protons. But the assumptions for the form of the field theory are difficult to justify for the large time-like momenta transfer involved. Further, one needs cross sections for excited baryons states, such as $\Delta$s and $\Omega$. The annihilation cross section for hyperons are also lower according to the additive quark model \cite{URQMD}. It is safe to state that annihilation cross sections could easily vary by several tens of percent, and that these variations would be very difficult to calculate from first principles. One can also use fits to heavy ion data, but in these cases one needs to first rely on an assumption for the initial production rates in dense matter. One might expect annihilation cross sections to be smaller for hyperons since hyperons have smaller charge radii than nucleons. For that reason, we consider a simple parameterization where the dependence of the annihilation cross section on strangeness is encapsulated in a single parameter, $\alpha$. We then repeated our calculations above with different values of $\alpha$. The form for the cross section for the annihilation of two baryons $A$ and $B$ with relative three-momentum $P$ is \begin{equation} \label{eq:sreduction} \sigma_{\bar{A}B}(Q)=\sigma_{\bar{p}p}(Q)\left\{\frac{1}{2} \left[\alpha^{|S_A|}+\alpha^{|S_B|}\right]\right\}^2. \end{equation} This expression describes a physical picture where the particles annihilate whenever the distance of closest approach is less than the sum of the two baryon's radii $R_A$ and $R_B$, and where the size of the baryon scales as \begin{equation} R_A=R_p\alpha^{|S_A|}. \end{equation} For $\alpha=1$, one reproduces the cross sections described in Eq. (\ref{eq:xsection}). For $\Omega\bar{\Omega}$ annihilation the cross section is reduced by a factor $\alpha^3$. Results for $\alpha=0.75$ are displayed in Fig. \ref{fig:poverpi}. Even though this is a rather strong dependence on the strangeness content, the change in the annihilation fractions are rather modest. Ratios for $\Xi/\pi$ and $\Omega/\pi$ ratios are also shown from ALICE \cite{Abelev:2013zaa}. The ALICE results are in the range of the predictions of the model that include baryon annihilation. The chemical rate calculations ignore $2\leftrightarrow 2$ processes that affect the hyperon/proton ratios. For example, $p+K\leftrightarrow \Lambda +\pi$ has a preference for moving strangeness from kaons into hyperons due to the fact that $\mu_p+\mu_K>\mu_\Lambda+\mu_\pi$. This would increase the $\Lambda/p$ ratio. In contrast, the process $K+\Lambda\leftrightarrow p+\pi$ prefers to reduce the number of strange quarks and reduce the $\Lambda/p$ ratio. Since these processes involve momentum transfers on the order of a few hundred MeV, they can probably be calculated with some confidence. In the URQMD calculations of \cite{Steinheimer:2012rd,Becattini:2012sq,Becattini:2012xb} the number of $\Omega$s and $\Lambda$s increased in the hadronic stage, which is in contrast to the results found here. It would be useful to further analyze the URQMD evolution to understand what is driving this difference. \section{Conclusions} Figure \ref{fig:alldens} shows that annihilation processes during the hadronic stage play a significant role in the baryo-chemistry of the quark-gluon plasma. Calculated baryon yields that include the effects of annihilation are reduced by $\sim 40\%$ for central Pb+Pb collisions at the LHC. This helps explain the magnitude of the $p$ and $\bar{p}$ spectra at the LHC relative to the pion spectra without surrendering the assumption of chemical equilibrium at the advent of the hadronic stage. Figure \ref{fig:alldens} also displays results from calculations where the inverse reactions, e.g. $5\pi\rightarrow p\bar{p}$, were not included. In this case baryon yields would have fallen approximately 50\% if regeneration had not been incorporated. This is consistent with microscopic transport models that also ignore regeneration \cite{Steinheimer:2012rd,Becattini:2012sq,Becattini:2012xb}. Regeneration increases this reduced yield by about 20\% so that the net reduction in baryon number is close to 40\%. This is in line with observations at the LHC, which have uncertainties at the 15\% level. Although it is clearly important to include annihilation, one may or may not wish to include regeneration depending on the level of accuracy required for a specific analysis. Certainly, $n\rightarrow 2$ processes can require significant costs, both in the development and the execution of the code. Regeneration perfectly cancels the annihilation rate immediately after chemical equilibrium, then becomes negligible by the end of the reaction. This implies that baryo-chemistry is most sensitive to the physics for temperatures $\lesssim 140$ MeV. For these temperatures, hadronic cascades are well justified. Given that the assumption of local kinetic equilibrium becomes increasingly questionable once the temperature falls below 140 MeV, the results found here are suspicious. However, the annihilation fraction without regeneration was compared to that of a microscopic simulation and found to match within one or two percent. Certainly, the effects of regeneration could vary a few percent between this simple model and a more realistic model. Thus, this simple calculation mainly serves as a means to gauge whether regeneration, which requires significant work to implement in a microscopic model, warrants the effort to incorporate into microscopic models. The answer is ``yes'', if one wants to understand baryon yields to the 10\% level or better, but ``no'' if 20\% accuracy is sufficient. It is also clear that baryon annihilation cannot be ignored even if 20\% accuracy is sufficient. Even after regeneration is included, the annihilation fraction is in the neighborhood of 40\%. Figure \ref{fig:alldens} also showed the effect of annihilation on strange baryon yields. However, these results are not trustworthy given the lack of knowledge of annihilation cross sections for hyperons. Whereas the nucleon-antinucleon cross sections do not strongly depend on isospin, e.g. $\bar{p}p$ and $\bar{p}n$ annihilation cross sections are similar \cite{pbarn}, little is known about the annihilation cross sections for hyperons, either with antinucleons or antihyperons, and one must resort to unconstrained model calculations \cite{kovacs} or to fits to particle yields in heavy ion collisions \cite{Wang:1998sh}. Additionally, it is possible that $2\rightarrow 2$ inelastic processes such as $K+p\rightarrow \pi+\Lambda$, which are not included here, might affect yields. In our calculations $\mu_K+\mu_p>\mu_\Lambda+\mu_\pi$, so such effects might alter the $\Lambda/p$ ratio. Any study of hyperon yields in heavy ion collisions will remain largely speculative until experimental information is obtained regarding hyperon annihilation reactions in the vacuum. The analysis of this paper, and of similar works, would significantly benefit if the details of how yields were measured was clearly provided. This is especially true for understanding the degree to which weak resonance decays are included in experimental results, especially preliminary results. For example, at the 10\% level ratios of baryon yields to pions depend on whether weak decays, e.g. $K_s\rightarrow \pi\pi$, that produce pions are included. Presentations of model results are often similarly vague in their description. All weak decays were consistently ignored in the calculations presented here, but that can easily be modified to better either match experimental conditions or to match the choices of a competing theoretical model. Although sufficient for obtaining an understanding of the importance of annihilation chemistry during the hadronic stage, the model employed here is too simplistic to compare to data at better than the 10\% level. The assumptions are numerous: hadronic equilibrium at the onset of hadronization, boost-invariant dynamics, and local kinetic equilibrium that persists for long times. Further, some of the chemistry is neglected. Inelastic collisions, such as $\Delta N\rightarrow NN$, can lower the final meson yields on the order of 10\%. As shown here, to make a hadronic cascade reliable to the 10\% level, one should employ a fully consistent microscopic simulation of the hadronic stage that includes regeneration. The cascade would then account for the three-dimensional expansion, include more $2\leftrightarrow 2$ processes, provide a realistic freeze-out picture, and allow for the loss of local kinetic equilibrium.
\section{Introduction} In first order formulations of general relativity one has a notion of local Lorentz invariance, which can be thought of as one way of implementing the equivalence principle \footnote{Linking my talk at this wonderful conference to Einstein's Prague days.}. It is crucial to understand the fate of this gauge symmetry in attempts to quantise gravity, both theoretically and with regard to a possible phenomenology of quantum gravity (including matter). There are strong experimental constraints on many possible types of violation of Lorentz covariance and any proposed theory of quantum gravity must prove itself consistent with such constraints. In Hamiltonian formulations, in particular the Ashtekar-Barbero connection formulation \cite{ashtekar,barbero}, the issue of Lorentz covariance has been the focus of some debate, since the Ashtekar-Barbero formulation naturally uses the gauge group $\SU(2)$ or $\SO(3)$\footnote{The covering group $\SU(2)$ is required if one wants to include spinors. We consider pure gravity; the symmetry groups we discuss arise as the isometry groups of real manifolds or the stabilisers of points in them, and can be taken to be real-valued matrix groups. By expressions such as $\SO(3,1)$, we mean the connected component preserving orientation and time orientation.}, instead of the full Lorentz group. The use of this smaller gauge group is connected to the appearance of second-class constraints in previous attempts to maintain full Lorentz covariance. Here we show how to avoid second class constraints and stay Lorentz covariant by introducing a field of {\em local observers}. Details are given in the paper \cite{lorentz}. \section{Canonical First Order General Relativity} Starting from the Lorentz covariant Palatini-Holst action for vacuum general relativity without cosmological constant \ben S[e,\omega]=\frac{1}{8\pi G}\int\kappa_{abcd}\,e^a\wedge e^b\wedge R^{cd}[\omega]\,, \label{palholst} \een where $\kappa_{abcd}$ is an $\SO(3,1)$-invariant bilinear form on $\mathfrak{so}(3,1)$, \ben \kappa_{abcd}=\half\epsilon_{abcd}+\frac{1}{2\gamma}(\eta_{ac}\eta_{bd}-\eta_{ad}\eta_{bc})\,, \een one can perform the usual canonical analysis and find that the 18 momenta $\pi_{ab}^i$ conjugate to the spatial components of the connection $\omega^{ab}_i$ are expressible in terms of only 12 tetrad components $e^a_i$. This leads to {\em second class constraints}, which provide an obstacle to quantisation and usually require introducing new variables which are harder to interpret in terms of spacetime geometry. In Holst's analysis \cite{holst} leading to the well-known Ashtekar-Barbero formulation of canonical gravity, one deals with this issue by explicit symmetry breaking to $\SO(3)$: Imposing `time gauge' $e^0_i=0$ and defining \ben A^{ab}=\omega^{ab}+\frac{\gamma}{2}{\epsilon^{ab}}_{cd}\omega^{cd}\,, \een only the $\mathfrak{so}(3)$ part of $A$ (the {\em Ashtekar-Barbero connection}) has nonvanishing conjugate momentum, and one avoids second class constraints. However, this comes at the price of losing Lorentz symmetry which is broken explicitly by the gauge choice. In our formalism we replace time gauge by a condition involving a field of {\em internal observers} $y$ which specifies a time direction locally, and leads to a spontaneous breaking of symmetry from $\SO(3,1)$ to a subgroup $\SO(3)_y$ depending on $y(x)$ at each spacetime point $x$. \section{General Relativity with Local Observers} For a given spacetime manifold with metric $g$ or frame field $e$, we define a {\em field of observers} as a unit future-directed timelike vector field $u$. Using the frame field we can map it to a spacetime scalar $y=e(u)$ valued in the velocity hyperboloid ${\rm H}^3=\SO(3,1)/\SO(3)$. But such a field of {\em internal observers} can be defined without specifying the metric, and is hence suitable for a framework in which the metric arises dynamically as a solution to the equations of motion. Our formalism for {\em generalised canonical gravity} builds on the following variables: \begin{itemize} \item a field of internal observers $y$, valued in ${\rm H}^3\subseteq \R^{3,1}$, thought of as giving a local notion of time direction, \item a nowhere-vanishing 1-form $\hat{u}$, thought of as non-dynamical and generalising the normal to a foliation (if $\hat{u}\wedge d\hat{u}=0$, $\hat{u}$ is of the form $\hat{u}=N\,dt$) -- one can always reduce to the case of a foliation by choosing an appropriate $\hat{u}$, \item an $\R^3_y$-valued `triad' 1-form $E$, where $\R^3_y$ is the subspace of $\R^{3,1}$ orthogonal to $y$ (this generalises time gauge). \end{itemize} The spacetime coframe field is then simply given by \ben e=E+\hat{u}\,y \label{one} \een analogous to how one reconstructs the spacetime metric in the ADM formulation using lapse and shift. As is usual in first order gravity, we must require $e$ to be nondegenerate. The field of internal observers $y$ defines a field of spacetime observers by $y=e(u)$, and one finds that $E(u)=0$ so that $E$ is actually {\em spatial}. Similarly, we define spatial and temporal parts of the spin connection, \ben \omega = \Omega + \hat{u}\,\Xi \label{two} \een Substituting (\ref{one}) and (\ref{two}) into the Palatini-Holst action (\ref{palholst}) gives us a generalised Hamiltonian formulation of vacuum general relativity in terms of an action depending on $y,E,\Omega$ and $\Xi$ that we give in \cite{lorentz}. Up to this stage everything is Lorentz covariant -- we have just changed variables in the action. The r\^{o}le of the field of internal observers $y$ is to give us a local embedding of $\SO(3)$ into $\SO(3,1)$. The embedding can be freely changed by applying a Lorentz transformation $y\mapsto y'=\Lambda\,y$; allowing those Lorentz transformations instead of thinking of $y$ as fixed restores Lorentz covariance. The spatial connection $\Omega$ can be projected to its $\mathfrak{so}(3)_y$ part ${\bf \Omega}$. Then under a local Lorentz transformation \ben {\bf \Omega} \mapsto {\bf \Omega}' = \Lambda^{-1}\,{\bf \Omega}\,\Lambda+\pi_{y'}(\Lambda^{-1}\,d^{\perp}\Lambda)\,, \een where $\pi_{y'}$ is a projector onto $\mathfrak{so}(3)_{y'}$ and $d^{\perp}=d-\hat{u}\wedge\pounds_u$ is a spatial exterior derivative. Therefore, if one only applies $\SO(3)_y$ transformations which leave $y$ invariant, ${\bf \Omega}$ transforms as an $\SO(3)_y$ connection, while if one allows for transformations that rotate the local internal observer $y$ to $y'$, the transformed connection ${\bf\Omega}'$ is in $\mathfrak{so}(3)_{y'}$. This is as it should be. To understand the dynamical structure of this formalism, we focus on the term in the action that determines the symplectic structure in Hamiltonian general relativity, \ben S=\frac{1}{8\pi G}\int\kappa_{abcd}\hat{u}\wedge E^a\wedge E^b\wedge \pounds_u\Omega^{cd}+\ldots \een Since $E\wedge E$ is valued only in $\mathfrak{so}(3)_y$, only half of the components of $\Omega$ have nonvanishing conjugate momentum. The number of independent components of $E$ matches the number of conjugate momenta, and no second-class constraints arise -- but we did not find it necessary to impose any gauge fixing such as the time gauge employed in Holst's analysis. One can make the splitting of $\mathfrak{so}(3,1)$ into a rotational subalgebra $\mathfrak{so}(3)_y$ and a complement $\mathfrak{p}_y$ explicit by choosing local bases $J^{ab}_I$ and $B^{ab}_I$ (depending on $y$). Then \ben A^I:={\bf \Omega}^I+\gamma K^I\,, \label{ashbarb} \een is conjugate to $(E\wedge E)^I$, where ${\bf \Omega}$ and $K$ are the $\mathfrak{so}(3)_y$ and $\mathfrak{p}_y$ parts of $\Omega$. (\ref{ashbarb}) is the Ashtekar-Barbero connection, and our formalism is dynamically equivalent to the Ashtekar-Barbero formulation: It has the same phase space variables, subject to the same constraints that define the dynamics. In the form (\ref{ashbarb}) manifest Lorentz covariance is lost; it can be recovered by viewing $\mathfrak{so}(3)_y$ and $\mathfrak{p}_y$ not as fixed (isomorphic) representations of $\SO(3)$, but as subspaces of $\mathfrak{so}(3,1)$ specified by the field $y$. \section{Cartan Geometrodynamics} Situations of spontaneous symmetry breaking in gravitational theories are geometrically best understood in terms of {\em Cartan geometry} \cite{derekproc}. A well-known example is the MacDowell-Mansouri formulation \cite{macdo} of gravity with cosmological constant (we take $\Lambda>0$ but $\Lambda<0$ is analogous) in terms of the $\SO(4,1)$ invariant action \ben S_{\rm MM} = -\frac{3}{32\pi G\Lambda}\int\epsilon_{abcde}\left(F^{ab}\wedge F^{cd}\right)y^e\,, \label{macdo} \een where $F$ is the curvature of an $\SO(4,1)$ connection $A$. The field $y$ takes values in de Sitter spacetime $\SO(4,1)/\SO(3,1)\subseteq \R^{4,1}$; it breaks the symmetry at each point in spacetime to the subgroup $\SO(3,1)_y$ leaving $y$ invariant. Fixing $y=(0,0,0,0,1)$ in the action breaks the symmetry explicitly. The Lie algebra $\mathfrak{so}(4,1)$ splits into a subalgebra $\mathfrak{so}(3,1)_y$ and a complement $\mathfrak{t}_y$; identifying the $\mathfrak{so}(3,1)_y$ part of $A$ with the spin connection $\omega$ and the $\mathfrak{t}_y$ part with a coframe $e$, \ben A = \left(\begin{array}{c c} \omega & \sqrt{\frac{\Lambda}{3}}e \\ -\sqrt{\frac{\Lambda}{3}}e & 0 \end{array}\right)\,, \label{cartanco} \een the action (\ref{macdo}) reduces to the Einstein-Hilbert-Palatini action with a cosmological term. Cartan geometry is about infinitesimally approximating the geometry of a curved manifold by a homogeneous spacetime $G/H$ (in this case de Sitter spacetime) which generalises the tangent space $\R^{p,q}$ used in (pseudo-)Riemannian geometry. The Cartan connection $A$ relates the model spacetimes tangent to different points of the manifold -- for a model spacetime of non-zero curvature, $A$ is flat if the manifold is (locally) isomorphic to the model spacetime. This naturally introduces a cosmological constant into gravity, given by the curvature scale of the model spacetime. Our reformulation of the Ashtekar-Barbero formalism for canonical gravity is best interpreted as describing the geometry of {\em space} as {\em Cartan geometrodynamics}: The $\mathfrak{so}(3)_y$ connection ${\bf \Omega}$ (or, alternatively, the Ashtekar-Barbero connection) and the triad $E$ can be assembled into a Cartan connection \ben {\bf A} = \left(\begin{array}{c c} {\bf \Omega} & \frac{1}{l}E \\ 0 & 0 \end{array}\right)\,, \label{cartangeo} \een taking values in the Lie algebra of the Euclidean group $\mathfrak{iso}(3)$ if we consider a vanishing cosmological constant ($l$ is an (unspecified) length scale put in for dimensional reasons). The appearance of the group $\ISO(3)$ is understood as follows: Spacetime is infinitesimally modelled on Minkowski spacetime, with isometry group $\ISO(3,1)$. At a given point in spacetime, picking an observer in the model Minkowski spacetime gives a notion of `space' in the model spacetime as the maximal totally geodesic hypersurface orthogonal to this observer -- in the construction above, we referred to this as the subspace $\R^3_y$ orthogonal to an observer $y$. This breaks the symmetry to $\ISO(3)$, the isometry group of $\R^3_y$. Picking a point in $\R^3_y$ tangent to the spacetime point then breaks the symmetry further to $\SO(3)$, giving the splitting (\ref{cartangeo}). For a more detailed discussion of the geometry behind Cartan geometrodynamics we refer to \cite{obssp}. \section{Summary and Outlook} We have given a reformulation of canonical general relativity in first order form which uses local observers that define a local notion of time. These give an embedding of the rotational subgroup $\SO(3)$ into the Lorentz group that allows to reconstruct Lorentz covariance from the $\SO(3)$ Ashtekar-Barbero formulation of canonical gravity. The geometry behind our constructions is best understood in terms of {\em Cartan geometrodynamics}. Since this formulation requires only a local choice of time direction not necessarily related to a foliation of spacetime, it links the canonical and covariant formulations of general relativity \cite{cancov}. It would be important to understand the coupling of matter -- which would be necessary to investigate the possibility of physically observable Lorentz violation -- and the role of the field of internal observers there. So far they have been treated like lapse and shift, as Lagrange multipliers. Making the observer field dynamical could relate our framework to models with dynamical reference frames, such as Brown-Kucha\v{r} dust \cite{brownkuch}. Similar constructions could also be useful in approaches to quantum gravity where local Lorentz covariance is not manifest, such as Ho\v{r}ava-Lifshitz gravity, shape dynamics or causal dynamical triangulations. Taking the idea of local observers one step further, it is natural to consider the space of all possible choices of local observer -- {\em observer space}. In general relativity, this is the direct product of spacetime with the local velocity space ${\rm H}^3$ of normalised future-directed timelike vectors, but we consider it as a seven-dimensional manifold in its own right and study its geometry, both in general relativity and in more general settings. This is the viewpoint adopted in the work \cite{obssp}, where we show how the Cartan connection $A$ specified by a frame field $e$ and a spin connection $\omega$ as in (\ref{cartanco}) gives a Cartan geometry on observer space, with model space $\SO(4,1)/\SO(3)$, the space of all observers in de Sitter spacetime. Conversely, we investigate integrability conditions that allow the reconstruction of an invariant spacetime starting from an observer space Cartan geometry (i.e. a general Cartan geometry modelled on $\SO(4,1)/\SO(3)$); intuitively, such a reconstruction is possible if the connection is flat in the `velocity' directions of observer space. Different approaches to quantum gravity and quantum-gravity phenomenology incorporate the idea that spacetime geometry is an observer-dependent (or `momentum-dependent'), relative concept. From the perspective of observer space, such ideas correspond to observer space Cartan connections that are not flat in velocity directions, so that no invariant spacetime can be reconstructed. One example is the proposal of relative locality \cite{relloc} which suggests that `spacetime' and hence the notion of locality are observer-dependent, but there is an invariant momentum space shared by all observers. In \cite{obssp} we find that the framework of relative locality corresponds to an observer space connection that is flat in `spacetime', not `velocity' directions. For a general observer space geometry, both `spacetime' and `velocity space' are only defined relative to an observer. It will be interesting to see whether other ideas, such as that of an `effective metric' $\langle g_{\mu\nu}\rangle_k$ (depending on a momentum scale $k$) that appears in the asymptotic safety scenario for quantum gravity \cite{asympt}, can be discussed in the framework of observer space geometry. {\bf Acknowledgements.} -- I would like to thank Derek Wise for collaboration on the papers \cite{obssp,cancov,lorentz} that discuss the ideas presented in this proceedings contribution in detail and for comments on the manuscript. Research at Perimeter Institute is supported by the Government of Canada through Industry Canada and by the Province of Ontario through the Ministry of Research \& Innovation. \section*{References}
\section{Introduction} Interacting and interdependent networks have recently {attracted} great attention \cite{Havlin1,Vespignani,Havlin2,Havlin3,Grassberger,Dorogovtsev,Yamir1, Yamir2,Jesus,Ivanov}. Here, the function of a node in one network depends on the operational level of the { nodes it is dependent on} in the other networks. Investigated examples {range} from infrastructure networks as the power-grid and the Internet \cite{Havlin1} to interacting biological networks in physiology \cite{Ivanov}. Understanding how critical phenomena \cite{crit, Dynamics} are affected by the presence of interaction{s} or interdependent networks is crucial to control and monitor the dynamics of and on complex systems. In this context it was shown that interdependent networks are more fragile than single networks, and that the percolation transitions can be first order \cite{Havlin1}. Interesting, but so far less explored, is the case of interacting social networks, describing individuals that manage their personal relationships in different social contexts (e.g., work, family, friendship, etc.). Taking into account these multiple layers is crucial, as proven recently for community detection methods in social networks \cite{Fortunato_com, Lambiotte, Lehmann}, but the effect of their presence is still not understood in many respects. For example, there is considerable current interest in opinion models \cite{Fortunato}, among which we cite the the Sznajd model \cite{Sznajd}, the voter model \cite{Redner}, the naming game \cite{Andrea,Korniss} and Galam models \cite{Galam,Galam_05}. {But the influence of more than one network has gathered less attention\cite{galam_percolation}.} Here we propose a simple model for opinion dynamics that describes two parties competing for votes during a political campaign. Every opinion, i.e., party, is modeled as a social network through which a contagion dynamics can take place. Individuals, on the other hand, are represented by a node on each network, and can be active only in one of the two networks (vote for one party) at the moment of the election. Each agent has also a third option \cite{Andrea,Lama,ng_opinion,Korniss,Strogatz}, namely not to vote, and in that case she will be inactive in both networks. Crucially, agents are affected by the opinion of their neighbors, and the nodes tend to be active in the networks where their neighbors are also active. Moreover, the chance of changing opinion decreases as the decision moment approaches, in line with the observation that vote preferences stabilize as the election day comes closer \cite{polls}. Our aim is to provide insights in the role of multiple social networks in the voting problem through a simple and clear mathematical model, in the spirit, for example, of recent work concerning the issue of ideological conflict \cite{Strogatz}. We describe the dynamics of social influence in the two networks, and we model the uncertainty reduction preceding the vote through a simulated annealing process. {Long} before the election the agents change opinions and can sustain a small fraction of antagonistic relations, but as the election approaches their dynamics slows down, until they reach the state in which the dynamics is frozen, at the election day. At that moment, the party winning the elections is the one with more active nodes. Finally, we focus on the case in which the networks sustaining each party are represented by two Poisson graphs, and address the role of different average connectivities. This choice is consistent for example with the data on social networks of mobile phone communication, {which} are characterized by a typical scale in the degree (being fitted with a power-law distribution of exponent $\gamma=8.4$) \cite{Onnela}. We observe a rich phase diagram of the opinion dynamics. The results are that in the thermodynamic limit the most connected network wins the election {independent of} the initial condition of the system, in agreement with recent results on the persuasive role of a densely connected social network \cite{centola}. However, for a large region of the parameters the voting results of the two parties are very close and small perturbations could alter the results. In this context, we observe that a small minority of committed agents can reverse the outcome of the election result, thus confirming the results obtained in very recent and different models \cite{Korniss,Strogatz}. \section{Parties as antagonistic social networks} We consider two antagonistic networks $A, B$ representing the social networks of two competing political parties. Each agent $i$ is represented in each network and can choose to be active in one of the networks. In particular $\sigma^A_i=0$ if agent $i$ is inactive in network $A$ and $\sigma_i^A=1$ if agent $i$ is active in network $A$. Similarly $\sigma_i^B=0,1$ indicates if a node is active or {inactive} in network $B$. Since ultimately the activity of an individual in a network corresponds to the agent voting {for} the corresponding party, each agent can be active only on one network on the election day (i.e. if $\sigma_i^A=1$ then $\sigma_i^B=0$ and if $\sigma_i^B=1$ then $\sigma_i^A=0$). Nevertheless we leave to the agent the freedom not to vote, in that case $\sigma_i^A=\sigma_i^B=0$. Moreover agents are influenced by their neighbors. Therefore, we assume that, {on} the election day, if at least one neighbor of agent $i$ is active in network $A$, the agent will be active in the same network (network A) \textit{provided that it is not already active in network $B$}. We assume that a symmetrical process is occurring for the opinion dynamics in network B. Hence, the mathematical constraints that our agent opinions need to satisfy at the election day are: \begin{eqnarray} \sigma_i^A&=&\left[1-\prod_{j\in N_A(i)} (1-\sigma_j^A)\right](1-\sigma_i^B)\nonumber \\ \sigma_i^B&=&\left[1-\prod_{j\in N_B(i)} (1-\sigma_j^B)\right](1-\sigma_i^A), \label{cons} \end{eqnarray} where $N_A(i)$ ($N_B(i)$) are the set of neighbors of node $i$ in network A (network B). Therefore at the election day people cannot anymore change their opinion. On the contrary before the election we allow for some conflicts in the system, and in general the constraints provided by Eqs.~$\ref{cons}$ will not be satisfied. \begin{figure} \begin{center} \includegraphics[width=0.6\columnwidth]{figure1.eps} \end{center} \caption{(Color online) The two competing political parties are represented by two networks. Each agent is represented in both networks but can {either be active (green node) in only one of the two or inactive (red node) in both networks.} Moreover the activity of neighbor nodes influences the opinion of any given node.} \label{fig1} \end{figure} \section{Evolution dynamics during the election campaign} To model how agents decide {on} their vote during the pre-election period we consider the following algorithm. We consider a Hamiltonian that counts the number of the constraints in Eq. (\ref{cons}) that are violated. Therefore we take a Hamiltonian $H$ {of} the following form \begin{eqnarray} H&=&\sum_i\left\{\sigma_i^A-\left[1-\prod_{j\in N_A(i)} (1-\sigma_j^A)\right](1-\sigma_i^B)\right\}^2+\nonumber \\ &&\sum_i\left\{\sigma_i^B-\left[1-\prod_{j\in N_B(i)} (1-\sigma_j^B)\right](1-\sigma_i^A)\right\}^2. \end{eqnarray} {The terms in the brackets can take on the values $\pm1, 0$, therefore a natural choice of Hamiltonian to count the number of constraint violations involves squares of these terms.} We start from {an} initial condition {where the active nodes in networks A and B are distributed uniformly randomly}, and we consider the fact that {long} before the election the agents are free to change opinion. Therefore we model their dynamics as a Monte Carlo dynamics which equilibrate{s} following the Hamiltonian $H$ {starting from} a relative{ly} high initial temperature, i.e. {initially} some conflicts are allowed in the system. {Therefore, initially the active nodes in networks A and B are distributed according to the high temperature Gibbs measure, mimicking an effectively ``unbiased'' population at the beginning of the campaigning process. Moreover we note here that since we start with a sufficiently high temperature, the dynamics is not affected by the specific initial conditions of the system.} As the election day approach{es}, the effective temperature of the opinion dynamics decreases and the agents tend to reduce to zero the number of conflicts with their neighbors. The opinion dynamics described in this way is implemented with a simulated annealing algorithm. We start at { a} temperature $T=1$ and we allow the system to equilibrate by $2N$ Monte Carlo steps where a node is picked randomly in either {one} of the networks with equal probability and {is changed from active to inactive or vice versa. Subsequently, the Hamiltonian, or the number of conflicts, is recalculated. If the opinion flip results in a smaller number of conflicts, it is accepted. Else, it is accepted with probability $e^{-\Delta H /kT}$}. {This Monte Carlo process is repeated by slowly reducing} the temperature by a multiplicative factor {of} $0.95$ until we reach the temperature state $T=0.01$ where the Hamiltonian is $H=0$, there are no more conflicts in the network, and the probability of one spin flip is about $e^{-1/0.01}\simeq 10^{-44}$. {The choice of increment in the temperature reduction is such that the overall simulation time is compatible with the dynamics of social systems. The Monte Carlo sweeps that are performed, each of which corresponds to one campaigning day, span a total number of $\log 0.01 / \log 0.95 \approx 90$ days.} It turns out that the Hamiltonian $H$ has in general multiple fundamental states and the simulated annealing algorithm always find one of these states. The final configuration for the model just described is depicted in Figure $\ref{fig1}$. In Figure $\ref{SA}$ we report the result of this opinion dynamics for two antagonistic networks {A, B} with Poisson degree distribution{s} and different average connectivities {$z_A$, $z_B $, respectively}. In particular we plot the size $S_A$ of the giant component of the percolating cluster in network A, {i.e. the largest connected component of active nodes in network A}. \begin{figure} \begin{center} \includegraphics[width=0.8\columnwidth, height=11cm]{figure2a.eps} \end{center} \caption{(Color online) (Panel A) The size of the largest connected component $S_A$ in network A at the end of the simulated annealing calculation as a function of the average connectivity of the two networks: $z_A$ and $z_B$ respectively. The data is simulated for two networks for $N=500$ nodes and averaged 60 times. The simulated annealing algorithm is independent {of} initial conditions. The white line represent the boundary between the region in which network A is percolating and the region in which network A is not percolating. (Panel B) The schematic representation of the different phases of the proposed model. In region I none of the networks is percolating, in region II network B is percolating, in region III network A is percolating, in region IV both networks are percolating. } \label{SA} \end{figure} Additionally we have characterized the finite size effects (see Figure $\ref{finite}$) and concluded that the phase diagram of the model is consistent with the following scenario valid in the limit of large network sizes: \begin{itemize} \item {\it Region (I): $z_A<1, z_B<1$ }. In this region both giant components in network A ($S_A$) and network B ($S_B$) are zero, $S_A=0, S_B=0$, and therefore essentially agents never vote. \item {\it Region (II) in Figure $\ref{SA}$: } In this region the giant component in network {B} emerges, $S_B>0, S_A=0$. \item {\it Region (III) in Figure $\ref{SA}$: } In this region the giant component in network {A} emerges, $S_A>0, S_B=0$. \item {\it Region (IV) in Figure $\ref{SA}$} In this region we have the pluralism solution of the opinion dynamics and both giant component in networks A and B are different from zero, $S_A>0, S_B>0$. \end{itemize} \begin{figure} \begin{center} \includegraphics[width=\columnwidth]{figure3.eps} \end{center} \caption{(Color online) We represent the fraction of nodes in the giant component $S_A$ of network A and in the giant component $S_B$ of network B in different regions of the phase space. In region II ($z_A=1.5, z_B=4$) the giant component in network A ($S_A$ ) disappear{s} in the thermodynamic limit while in region IV ($z_A=2.5, z_B=4$) it remains constant. The giant component in network B remains constant in the thermodynamic limit both in region II and in region IV. Each data point is simulated for the two networks for $N$ nodes and averaged 200 times.} \label{finite} \end{figure} In Regions II (III) the active agents in party B (party A) percolate the system while agents in party A (party B) remain concentrated in disconnected clusters. Nevertheless, if the average connectivity of the two antagonistic parties is comparable (Region IV), the system can sustain an effective pluralism of opinions with both parties percolating in the system. Therefore, we find the interesting result that if the connectivity of the two parties is large enough,i.e. we are in region IV of the phase diagram (Figure $\ref{SA}B$) the pluralism can be preserved in the model and there will be two parties with {a } high number of votes. In order for a party to win the election, it is necessary that the active agents percolate in the corresponding network. The election outcome, nevertheless, depends crucially on the total number of votes in network A, $m_A$ and the total number of votes in network B, $m_B$. In Figure $\ref{majority}$ we plot the difference between the number of votes in network A and the number of votes in network B. Very interestingly, we observe that the more connected party (network) has the majority of the votes. It is also worth noting that the final outcome of the election does not depend on the initial conditions. Overall, this result support{s} the intuition that if a party has a supporting network that is more connected it will win the elections, and is coherent with recent results concerning the role of densely connected social networks on the adoption of a behavior \cite{centola}. \section{Committed agents} Different opinion-dynamics models have recently considered the role of committed agents \cite{Strogatz,Korniss,galam_committed}. Here we explore the effect of committed individuals during the election campaign by considering a situation in which a fraction of the nodes {always remain} active in one of the two networks, never {changing their} opinion. Figure \ref{committed} shows that in Region IV a small fraction of agents $f\simeq 0.1 $ \textit{in the less connected network} can reverse the outcome of the election. Indeed the probability distribution $P=P(m_A-m_B)$ in different realization of the dynamics is shifted towards the party supported by the committed minority. Remarkably, this finding fits perfectly with the results of the radically different models proposed in \cite{Strogatz,Korniss}, and generalizes them to the case of political elections. The crucial role potentially played by committed minorities is thus suggested by different models in different aspects of social dynamics, suggesting the need for future work exploring these findings. \begin{figure} \begin{center} \includegraphics[width=0.8\columnwidth, height=6.0cm ]{figure4.eps} \end{center} \caption{(Color online) The contour plot for the difference between the total number of votes $m_A$ in party A (total number of agents active in network A) {and} the total number of votes $m_B$ in party B (total number of agents active in network B). The data is simulated for two networks for $N=500$ nodes and averaged 90 times. {It is clear that the larger the difference in average connectivity of the two networks, the larger the advantage of the more connected political party}. } \label{majority} \end{figure} \begin{figure} \begin{center} \includegraphics[width=\columnwidth]{figure5.eps} \end{center} \caption{(Color online) We represent the role of a fraction $f$ of committed agents in reverting the outcome of the election. In particular we plot the histogram of the difference between the {fraction of agents $m_B/N$ voting for party $B$ and the fraction of agents $m_A/N$ voting for party $A$} for a fraction $f_A$ of committed agents to party A, with $f_A=0$ and $f_A=0.1$ and average connectivities of the networks $z_A=2.5, z_B=4$. The histogram is performed for 1000 realizations of two networks of size $N=1000$. In the inset we show the average {number} of agents in network A ($m_A$) and agents in network B ($m_B$) as a function of the fraction of committed agents $f_A$. A small fraction of agents ($f_A\simeq 0.1$) is sufficient to reverse the outcome of the elections. The data in the inset is simulated for two networks for $N=1000$ nodes and averaged 10 times.} \label{committed} \end{figure} \section{Conclusions} In conclusion, we have put forth a simple model for the opinion dynamics taking place during an election campaign. We have modeled parties (or opinions) in terms of a social networks, and individuals in terms of nodes belonging to these social networks and connecting them. We have considered the case of antagonistic agents who have to decide for a single party, or for none of them. We have described the quenching of the opinions preceding the voting moment as a simulated annealing process where the temperature is progressively lowered till the voting moment, when the individuals minimize the number of conflicts with their neighbors. We have shown that there is a wide region in the phase diagram where two antagonistic parties survive gathering a finite fraction of the votes, and therefore {the existence of} pluralism in the election system. Moreover, we have pointed out that a key quantity to get a finite share of the overall votes is the connectivity of the networks corresponding to different parties. Nevertheless connectivity is not sufficient to win the elections, since a small fraction of committed agents is sufficient to invert the results of the voting {process}. Though deliberately basic, the model provides insights into different aspects of the election dynamics. {Moreover, from a broader perspective, our work proposes a general framework for the description of any opinion formation process involving different contexts/networks, where opinions are frozen at some point in time, and where the agents' behavior reflects the approach of that point such that they are initially less susceptible to influence from their neighborhoods (high initial temperatures) and attempt to reduce the level of frustration/conflict more strongly later (low temperatures).} In {future works} we plan to generalize the model by studying antagonistic networks with different topolog{ies}, such as competing scale-free and Poisson networks or two competing scale-free networks. Other extensions of this mode{l} could describe several competing parties, consider a threshold dynamics as the one triggering the opinion formation of the agents in \cite{centola}, or relax the hypothesis of purely antagonistic interactions, thus allowing the agents to express multiple preferences in a multi-layered opinion space.
\section{Introduction} Graphene, strictly two dimensional allotrope of carbon atom with its unique mechanical\cite{mechanical}, structural\cite{structural}, electronic\cite{electronic1,electronic2} and thermal properties,\cite{thermal} has been considered as a promising candidate for next generation electronic devices and numerous nanoscale applications. Ingenious methods have been proposed for its production.\cite{prod1,prod2,dai,iijima,tongay,taner} Intensive studies have been also carried out for controlling and modifying various properties of bare graphene. The adsorption of foreign atoms or molecules on bare graphene surface has been considered as an efficient method to attain this objective. Graphene oxide (GOX) is an example\cite{gox1,gox2} to show how the properties of graphene can be changed dramatically upon the adsorption of oxygen atoms. GOX is obtained through oxidative exfoliation of graphite, which can be visualized as an individual sheet of graphene decorated with epoxy (C-O-C) and hydroxyl (C-OH) groups on both sides and edges. Incidentally, GOX has been also an attractive material for large scale graphene production\cite{aksay} due to low-cost, simple and high yield reduction methods. Unfortunately, despite the oxidized graphite is a known material since last 150 years\cite{1800s} and great deal of experimental and theoretical research carried out recently,\cite{li,boukhalov_2008,yan_2009,yan-prb,akbar,wang_2010,xiang,sun,mattson,kim_nature} a thorough understanding regarding the interaction of oxygen with graphene and relevant reactions are not available yet due their stochastic nature. \begin{figure} \includegraphics[width=8cm]{Figure1.png} \caption{Various critical sites of adsorption on the 2D honeycomb structure of graphene and an oxygen atom adsorbed on the bridge site, which is found to be as the energetically most favorable site. Carbon and oxygen atoms are shown by gray and red balls, respectively. (b) Variation of energy of oxygen adatom adsorbed to graphene along $H \rightarrow T \rightarrow B$ directions of the hexagon. The diffusion path of a single oxygen adatom is shown by stars. (c) Charge density isosurfaces, isovalues and contour plots of oxygen adsorbed graphene in a plane passing through C-O-C atoms. (d) Same as (c) on the lateral plane of honeycomb structure. (e) Total and partial density of states projected to carbon and oxygen atoms. Calculations are carried out for supercell presented in (c) where O-O interaction is significantly small.} \label{Figure1} \end{figure} To understand the experimental data, various structural configurations of GOX have been proposed based on first principles calculations. Performing the analysis of various coverage models, Boukhalov \textit{et al.}\cite{boukhalov_2008} revealed that 100\% coverage of GOX is energetically less favorable than 75\% coverage. Also, while a coverage less than 25\% of GOX contains only hydroxyl groups, the mixed GOX consisting of both oxygen and hydroxyl is favored for higher coverage. In a later study, Yan \textit{et al}.\cite{yan_2009} suggested that it is energetically favorable for the epoxy and hydroxy groups to aggregate together to form specific types of strips with $sp^2$ carbon regions in between. In contrast, Wang \textit{et al.}\cite{wang_2010} argued that thermodynamically stable structures are fully covered without any $sp^2$ carbon. The domains of graphene monoxide with $N_{O}/N_{C}=$1 (\textit{i.e.} the ratio of number of oxygen $N_O$ to the number of carbon atoms $N_C$) is attained by the oxidation of both sides.\cite{mattson} Very recent study\cite{kim_nature} combining experimental results and first principles calculations shows that multilayer GOX is metastable at room temperature undergoing modifications and reduction with a relaxation time of approximately 35 days. At the quasi-equilibrium, the nearly stable oxygen coverage was reported as $\Theta$=0.38 and presence of C-H species is found to favor the reduction of epoxides and to a lesser extent hydroxyl groups with the formation and release of water molecules.\cite{kim_nature} From our point of view, there exists still controversies between theory and experiment. For example, yet the distribution of hydroxy and epoxy groups on GOX surface together with the trends related with their clustering or uniform coverage are unknown. At least, a rigorous explanation for the reason of the differences in the interpretations of experimental data is required. In particular, it is not clear why the desorption of oxygen adatoms through O$_2$ formation does not occur so easily despite the negative formation energy of oxygen adsorption. Unlike GOX, the hydrogenated graphene, i.e. graphane (CH)\cite{CH1,CH2} and fluorinated graphene, i.e. fluorographene (CF)\cite{CF1,CF2} are experimentally realized and their crystal structure are well understood. In this study we present an extensive analysis of the oxygen adsorption and oxygen coverage by using first principles calculations based on Density Functional Theory (DFT). In order to understand the reversible oxidation-deoxidation processes\cite{science,gox2} we consider only oxygen adatoms on graphene surfaces, in spite of the fact that hydroxyl groups are readily coadsorbed. Earlier studies have followed approaches, which consider the optimized geometries corresponding to the minimum of total energy. Here, we show that the mechanism of oxygen coverage is governed mainly by the charge density profile of graphene, which is modified by each adsorbed oxygen in the course of oxidation. At the end, oxidized regions of graphene tend to form domains instead of a uniform coverage. In view of these results we also discuss unzipping process of graphene.\cite{aksay,sun} The oxygen adsorption on both sides of graphene was shown to be energetically more favorable than the adsorption to only one side, whereby serious distortions of the graphene lattice occurred. The repulsive interaction between two oxygen adatoms at the close proximity is repulsive and hinders oxygen desorption through O$_2$ formation. We finally showed that the distribution of oxygen atoms on graphene affects the electronic properties. Even if the massless Dirac-fermion behavior of graphene can be recovered for patterns conserving specific symmetries, the band gap normally increases with increasing non-uniform oxygen coverage and attains the value as high as 3 eV. These results are critical for the device applications based on reversible oxidation-deoxidation of graphene surfaces.\cite{science,gox2} \section{Method} Calculations are carried out within spin-polarized and spin-unpolarized density-functional theory (DFT) using projector augmented wave (PAW) potentials.\cite{paw} The numerical calculations have been performed by using VASP package.\cite{vasp1,vasp2} The exchange correlation potential is approximated by generalized gradient approximation functional of Perdew, Burke, and Ernzerhof (PBE).\cite{pbe} Calculations are carried out using periodically repeating supercell geometry, where the spacings between graphene layers are taken 15 \AA. However, systems involving very large graphene sheets are treated with 10 \AA~spacing, which is still large and hinders interlayer coupling. A plane-wave basis set with kinetic energy cutoff of 500 eV is used. All atomic positions and lattice constants are optimized by using the conjugate gradient method, where the total energy and atomic forces are minimized. The convergence for energy is chosen as 10$^{-5}$ eV between two steps. Oxygen-adatom and graphene system breaks inversion symmetry and a net electric-dipole moment is generated perpendicular to the graphene surface. Dipole corrections\cite{dipole} are applied in order to remove spurious dipole interactions between periodic images. The $\Gamma$-point i.e. \textbf{k}=0 is used for rectangular supercells containing 128 carbon atoms and oxygen adatoms, while 18x18x1 \textbf{k}-point sampling is used for primitive unit-cell. The Gaussian smearing with a width of 0.1 eV is used in the occupation of electronic energy bands. \section{ Interaction of Oxygen Atom with Graphene} A thorough analysis of the interaction of single O atom with graphene is essential to understand the oxidation process. Here the adsorption of single (isolated) oxygen on graphene is represented using large supercells, where O-O interaction is minimized. Owing to its hexagonal crystal structure, there are three major sites for foreign atom adsorption on graphene as shown in Fig. \ref{Figure1} (a). The hollow (H) site is above the center of hexagonal rings formed by carbon atoms. The top (T) site lies on top of the carbon atoms and the bridge (B) site is above the middle of each bonds connecting two adjacent carbon atoms. The bridge site is found to be most favorable adsorption site for an oxygen atom. Earlier LDA calculations predicted also B-site as energetically favorable site.\cite{a-nakada} The variation of the total energy along H $\rightarrow$ T $\rightarrow$ B sites is presented in Fig. \ref{Figure1} (b). The energy barrier is 0.6 eV for an O atom diffusing from bridge to top site and the energy difference between bridge and hollow site is as high as 2.56 eV. Therefore, the migration paths of oxygen adatom with minimum energy barrier follow the honeycomb structure over the C-C bonds by going from B- to T-sites as illustrated by inset in Fig. \ref{Figure1} (b). On the other hand, the energy barrier against the penetration of an oxygen adatom from one side of graphene to the other side is as high as 6 eV.\cite{coating} This high energy barrier suggests that graphene can be used an ideal coating preventing surfaces from oxidation. We note that the hollow site of graphene is more favorable for other atoms\cite{can_li,haldun_tm} such as Li or Ti, while H and F atoms prefers the top site for adsorption.\cite{CH2,CF2} The binding energy of oxygen on graphene is defined as \begin{equation}\label{equ:binding} E_b=E_{T}[Gr] + E_{T}[O] - E_{T}[Gr+O] \end{equation} where $E_{T}[Gr+O],E_{T}[Gr],E_{T}[O]$ denote the optimized total energies of graphene with adsorbed oxygen, pristine graphene and free O atom, respectively. Our calculations show that $E_b = 2.35$ eV for the (2x2) graphene supercell containing 8 carbon and one oxygen atom, but it increases to 2.40 for (3x3) and to 2.43 for (4x4) supercells. For supercells larger than (4x4), which correspond to smaller oxygen coverage and hinders O-O coupling, the binding energy does not change and mimics the binding energy of single, isolated oxygen attached to graphene surface. The calculated binding energy for full coverage $\Theta$=0.5 (namely the ratio, $N_C$, $N_{O}/N_{C}$=0.5) is 2.80 (3.34) eV per oxygen atom for one-sided (two-sided) adsorption. The binding energy of single oxygen adatom increasing from 2.43 eV to 2.80 eV at full coverage indicates a significant O-O coupling. We note that the formation energy $E_{f}=E_{b}-E_{b,O_{2}}/2$, where $E_{b,O_{2}}$ is the binding energy of O$_2$ molecule) is negative for one-sided coverage indicates instability. However, this situation does not impose desorption of O through O$_2$ formation for reasons discussed in Sec. V. According to the Pauling scale, oxygen has an electronegativity of 3.44, which is the second highest in periodic table after fluorine (3.98) and hence the oxidation of graphene is expected to result in significant charge transfer between oxygen and carbon atoms. Our calculations using Bader analysis\cite{bader} estimates a charge transfer of 0.79 electrons from carbon atoms of graphene to oxygen. This charge is mainly transferred from the nearest two carbon atoms forming the bond above which the oxygen adsorbed at bridge site, while some nearby oxygen atoms also contribute to the charge transfer. Figure \ref{Figure1} (c) shows isosurface and isovalue (contour) plots of total charge for a plane passing through carbon atoms and oxygen. The direction of electron density increasing from carbon atoms towards oxygen atom is a clear indication of charge transfer. In addition, two carbon atoms below oxygen are slightly raised from the plane of other carbon atoms and the charge distribution of the bond between them is disturbed. In Fig. \ref{Figure1} (d) bird's-eye view of isosurface and isovalue of charge density profile of a single oxygen adsorbed to each (5x5) supercell of graphene is presented. The structure is symmetric and the oxygen atom is at the center. We label some of the bonds corresponding to nearby bridge sites as B1,B2,B3,B4 and denote their equivalent sites by primes. While the isosurface plots of all C-C bonds of bare graphene are identical, the adsorption of single oxygen modifies the charge distribution at its close proximity. In fact, the isosurfaces plotted for 0.3 (electron/\AA$^{3}$ )\cite{vesta} show that the electron population of specific bonds are higher. As clearly seen from the figure, B2 and B3 bonds contain more electrons than in B1 and B4 bonds. The reason for the electron depletion in these bonds is related with the donation of electrons from these bonds to adsorbed oxygen. Interestingly, the bonds B2, B3 and their four images contain more electron density compared to B1 and other bonds further away from the oxygen atom. For a better illustration of bond charge alternation, we also presented the isovalue plot of total charge density in the upper triangle of Fig. \ref{Figure1} (d). Again, the more isolines in the isovalue map corresponds to more charge at B2' and B3' compared to B1' and B4'. In this context, we note that the long range interactions and Friedel oscillations found in 1D carbon chain\cite{longrange} and 2D graphene\cite{bacsi} induced by adatoms. Finally we include the density of states (DOS) plot in Fig. \ref{Figure1} (e) for the system presented in Fig. \ref{Figure1} (a). The overall total DOS represents a profile similar to bare graphene DOS making a dip near the Fermi level corresponding to Dirac points and DOS projected to oxygen atom is represented by a peak around -2.5 eV. Later we show that the electronic density and band gap will change with oxygen coverage. \subsection{ Interaction of Single Oxygen Molecule ($O_2$) with Graphene} In contrast to oxygen atom, an oxygen molecule has a weak binding with graphene. We calculated its binding energy to be 115 meV which consists of 57 meV Van der Waals interaction\cite{grimme} and 58 meV chemical interaction. Its magnetic moment is 1.90 $\mu_B$, slightly smaller than the magnetic moment of free O$_2$ due to weak chemical interaction. The $O_2$ molecule lies parallel, approximately 3 \AA{} above the graphene plane, and do not induce any distortions to graphene honeycomb structure as in the case of single oxygen adsorption. Accordingly, the binding of O$_2$ to graphene is specified as physisorption. It is therefore concluded that graphene cannot be oxidized directly by O$_2$ molecule unless its dissociation into oxygen atoms takes places at the vacancy site.\cite{coating} \section{Coverage of graphene surface with oxygen atoms} \subsection{Coverage of oxygen on one side} \begin{figure*} \includegraphics[width=17.5cm]{Figure2.png} \caption{(Color Online) (a) Charge density isosurfaces in a rectangular supercell containing 128 carbon atoms and a single adsorbed O atom shown by a red dot. B$_{i,j}$ identifies a specific C-C bond, where $i$ indicates the total number of oxygen atoms in the supercell and $j$ labels some of the bond around adsorbed oxygen atom(s). (b)-(c) and (d) are same as (a), except that 2,3, and 4 oxygen atoms are adsorbed to the sites, which are most favorable energetically. (e) Energetically favorable configurations up to 11 oxygen atoms adsorbed on graphene deduced from charge. (f) Energetically favorable configuration (left) and less energetic, ordered configuration (right) for 12 O atoms. (g) Variation of the binding energies of the last oxygen adatom up to 12.} \label{Figure2} \end{figure*} Starting from single oxygen adatom, we next consider the adsorption of more oxygen atoms one at a time on graphene surface leading to higher coverage of oxygen. We exclude the hydroxyl groups in the present study to simplify the situation and hence to reveal essential aspects of oxygen adsorption. In order to reduce the effects of cell size, we construct a larger rectangular supercell containing 128 carbon atoms as shown in Fig. \ref{Figure2} (a). The isosurface charge density profile for rectangular supercell is similar to the charge density profiles in Fig. \ref{Figure1} (d) with $B_{1-2}$ and $B_{1-3}$ bonds having more charge compared to $B_{1-1}$ and $B_{1-4}$. For the adsorption of second oxygen, we try all inequivalent bridge sites and calculate their total energies. It turns out that, the energetically most favorable site for the second oxygen adsorption is at $B_{1-2}$ site in Fig. \ref{Figure2} (a). In addition, the calculated binding energy of the second oxygen is around 2.9 eV and this is even higher than the binding energy of single oxygen on graphene. The binding energy is 154 meV lower for adsorption on $B_{1-3}$. The binding energy at $B_{1-5}$ site is equal to the single oxygen binding energy. But interestingly, $B_{1-1}$ and $B_{1-4}$ sites are energetically less favorable sites for second oxygen adsorption compared to other sites. The calculated $E_{b}$ is 2.33 eV for $B_{1-1}$ site. These calculated energies indicate a direct correlation between the binding energies and isosurface profiles given in Fig. \ref{Figure2} (a). Apparently an oxygen atom prefers the bridge sites, where the electron density is highest compared to other available sites. The charge density isosurface profile of graphene supercell containing two oxygen atoms is presented in Fig. \ref{Figure2} (b) and this profile can be used to predict the energetically favorable and unfavorable sites for the adsorption of third oxygen. Again, there are some bridge sites such as $B_{2-1}$ and $B_{2-2}$ containing more electronic charge compared to other bonds like $B_{2-3}$ and $B_{2-4}$. The third oxygen is bound to $B_{2-1}$ site with $E_{b} = 3.06$ eV which is slightly higher for the maximum binding energy of second oxygen. The binding energies at $B_{2-3}$ and $B_{2-4}$ are approximately 0.7 eV smaller than the binding energy at $B_{2-1}$ site. For the case of fourth oxygen, $B_{3-2}$ site in Fig. \ref{Figure2} (c) having more bond charge compared to other sites is energetically most favorable. It's binding energy, $E_{b} = 2.90$ eV, is slightly smaller than the binding energy of previous oxygens. The favorable binding energy for fifth oxygen can be predicted as $B_{4-1}$ from Fig. \ref{Figure2} (c). The oxidation process of graphene for more than four oxygen is presented in Fig. \ref{Figure2} (e) and (f) up to 12 oxygens adatoms. The main trend is that each oxygen added to system prefers the bridge sites containing higher bond charge. For the sake of comparison, we included the ordered configuration for 12 adatoms in Fig. \ref{Figure2} (f). However, this configuration (right) is significantly less energetic, by 1.46 eV compared to to the random configuration on the left. We continue to examine the growth of the domain consisting of 12 atoms by adding oxygen atoms to the system. The 13$^{th}$ oxygen inserted to the system (not shown in figure) prefers the bridge site on the bond having highest electronic charge, but not the third bridge sites stacking eventually three oxygen atoms along a line of bridge sites of consecutive parallel C-C bonds. There are two such possible sites in Fig. \ref{Figure2}, which are identified as the precursors of unzipping (where the usual angle of C-O-C bridge bond increases by breaking (or weakening) the C-C bond underneath,\cite{aksay,sun} are energetically unfavorable by $\sim$0.9 eV. The 14th oxygen atom behaved like the previous one: instead of occupying two possible sites of precursor states, it is adsorbed to a different bridge site which is energetically 614 meV more favorable. Clearly, the final structure is a domain of oxygen adatoms on graphene for $\Theta <$ 0.5 and hence it lacks the signatures of any uniform coverage which is present for the case of hydrogen and flourine adsorption on graphene. In the case of oxygen, adatoms arrange themselves on graphene starting from a single adatoms. Subsequently, additional ones seek energetically most favorable sites clustering around the existing ones. This domain structure The binding energies of the last adsorbed oxygen atom (or $n^{th}$ adsorbed oxygen) is calculated from the expression $E_{b}[n] = (E_{T}[n-1] + E_{T}[O]) - E_{T}[n]$ in terms of the minimum total energies of $n-1$ and $n$ oxygen atoms adsorbed on the same supercell of graphene. For any $n$, the lowest total energy (with negative sign) $E_{T}[n]$ and hence highest binding energy (with positive sign) $E_{b}[n]$ is determined by comparing the calculated energies of $n^{th}$ oxygen adatom when adsorbed to sites remained from the domain of adsorbed $(n-1)$ oxygen atoms. Once oxygen adatoms nucleate a domain they prefer to grow it by including additional oxygen atoms, whereby uniform oxidation of graphene surface is precluded. This conclusion is attained by calculating the total energy of a single domain consisting of 12 oxygen atoms formed on a large graphene supercell consisting 256 carbon atoms and comparing it with the total energy of two separate domains of 6 oxygen atoms nucleated at two different locations on the same supercell. The growth of a single domain is found to be favored by 330 meV compared to the growth two separate domains. It should be noted that the present analysis is done under the condition, where sequential adsorption of oxygen adatoms achieved in equilibrium. However, oxidation is a stochastic process comprising processes or events taking place in nonequilibrium. Therefore, growth of multiple domains at finite temperature may occur, but the uniform growth appears to be a case of least probability. In Fig. \ref{Figure2} (g), the calculated binding energy, $E_{b}[n]$ exhibits an oscillatory variation for $n >2$ and $E_{b}[n=12]$=3.1 eV. These oscillations are physical since energetically favorable site for the $n$th adsorbed oxygen and the resulting $E_{b}$ are well-determined and unique, but not equivalent to previous sites. The oscillations of $E_{b}$ originate from the changes of the charge distribution and distortions of C-C bonds on the graphene surface occurred as a result of adsorbed oxygen atoms forming a domain. At this point, we note that the energetic sites of two and three oxygen adatom in Fig. \ref{Figure2} is in agreement with the adsorption sites found by Yan and Chou\cite{yan-prb} as well as by Sun and Fabris.\cite{sun} However, the ground state configuration of freely adsorbed oxygen atoms forming a domain comprising four or more atoms in Fig. \ref{Figure2} are different from that leading to the C-C unzipping, since the latter configuration first require to overcome a significant energy barrier.\cite{sun} As a matter of fact, two oxygen adsorbed to the bridge sites on two parallel C-C bonds of the same hexagon was unfavorable energetically by 1.5 eV.\cite{yan-prb,sun} On the other hand, at advanced stages of Fig. \ref{Figure2} comprising a domain of 8 oxygen atoms we obtained a configuration consisting two B-sites occupied by oxygen atoms on the adjacent parallel C-C bonds, which can be a precursor of unzipping if subsequently adsorbed oxygen atoms occupy additional adjacent B-sites on a line and they overcome an energy barrier to increase C-O-C bond by breaking the C-C bond underneath. However, next adsorbed oxygen as well as 13$^{th}$ and 14$^{th}$ stopped to develop the precursor state and preferred different sites. We note that at finite temperature and in the presence of other external effects oxygen atoms are prone to develop precursors of unzipping by deviating fro above sequence of adsorption taking place in equilibrium. Later in Sec. VII we discuss this issue further. We also note that patterns of oxygen coverage for $N_{O}/N_{C}$=0.5 predicted by the genetic algorithm\cite{xiang} cannot be directly comparable with the nonuniform coverage in the present study, which determines the most energetic sites when oxygen atoms are adsorbed on graphene sequentially one at a time. \subsection{High temperature behavior} In order to test the stability of oxygen covered region in Fig. \ref{Figure2} (f), we also performed finite temperature, ab-initio molecular dynamics calculations. The Nose thermostat is used and the time steps are taken 2.5 femtoseconds. Atomic velocities are normalized at every 40 time steps and calculations lasted for 10 picoseconds at 1000 K for supercell containing 128 carbon atom and 10 oxygen adatoms. At the end, this structure remained stable and neither $O_{2}$ formation, nor dissociation of oxygen atoms from graphene surface did occurred. High binding energies of adsorbed oxygen atoms and absence of oxygen dissociation or any irreversible structural transition at 1000 K suggest that the oxygen covered domains shall remain intact for reasonable times in spite of the fact that underlying graphene is locally distorted. \subsection{Coverage of oxygen on both sides of graphene } \begin{figure} \includegraphics[width=7cm]{Figure_altust.png} \caption{(Color Online) Adsorption of two oxygen atoms on one surface of graphene with buckling of 0.88 \AA. (b) Adsorption of two oxygen atoms at both sides with a buckling of 0.25 \AA. (c) Isosurfaces of bond charge densities after the adsorption of two oxygen atoms, each one adsorbed to different sides of graphene. Some of the C-C bonds of graphene, which are deprived of regular charges upon oxidation, are highlighted with arrows.} \label{Figure_altust} \end{figure} The adsorption of oxygen atoms on one side of graphene and hence formation of domain structure induces structural deformations at the underlying graphene. Fig. \ref{Figure_altust} (a) shows the side-view of graphene structure with two oxygen atoms adsorbed on the same surface of graphene. The carbon atoms below the adsorbed oxygen atoms are distorted and raised towards oxygen atoms. The resulting buckling is as large as 0.88 \AA{} above the plane of graphene. The amount of these distortions are further increased with the addition of more oxygen atoms. In contrast to this situation, the amount of buckling is reduced to 0.25 \AA, if the second oxygen atom were adsorbed to the other surface of graphene, as shown in Fig. \ref{Figure_altust} (b). Nonetheless, the latter configuration is 110 meV more energetic than the previous case. This result is good in agreement with Ref[\onlinecite{yan-prb}]. Hence, the two-sided adsorption shall be preferred instead of the single-sided adsorption. Despite that the favorable binding site of second oxygen atom at the other side follows our bond charge density analysis discussed in Sec. IV (B). For example, when an oxygen atom is adsorbed on one side as in Fig. \ref{Figure2} (a), the most favorable adsorption side for second oxygen is the same and is $B_{1-2}$ site no matter whether the second oxygen adsorbs to the top surface (one-sided adsorption) or to the bottom surface (two-sided adsorption). Moreover, the isosurface charge density profiles shown in Fig. \ref{Figure_altust} (c) for the case of second oxygen adsorbed on other side are identical to the profile in Fig. \ref{Figure2} (b) when two of the oxygens are adsorbed on one side. The ordering of energetically favorable sites presented in Fig. \ref{Figure2} for higher oxygen coverage is independent of the adsorption side. Nonetheless, two sided adsorption is energetically more favorable. \subsection{ Carbon (C) and Fluorine (F) adsorption on graphene } We now investigate the adsorption of carbon (C) and fluorine (F) atoms on graphene in the context of previous charge density analysis. Similar to oxygen atom, carbon atom is adsorbed at the bridge site. The atomic structure of carbon atom adsorbed on graphene presented in Fig. \ref{Figure_CF} (a) is reminiscent of the oxygen atom adsorption on graphene as presented in Fig. \ref{Figure1} (a). The distance between the adsorbed carbon atom and nearest-neighbor carbon atoms of graphene is 1.52 \AA, which is slightly larger than the distance of nearest carbon-oxygen atoms (1.46 \AA) in GOX. The angle formed between adsorbates and host graphene atoms which is 62.5 degree is almost equal for carbon and oxygen adsorption. On the other hand, the binding energy of carbon atom adsorbed on (5x5) supercell of graphene is 1.56 eV and significantly smaller than the binding of oxygen adatom on graphene. The Bader analysis calculates a charge transfer of 0.04 electrons from the adsorbed carbon atom at the bridge site to the host graphene atoms and this value is also significantly smaller and is in the reverse direction as compared charge transfer between carbon and oxygen atoms. Consequently, the chemical interaction between carbon atoms is covalent rather than ionic. Nonetheless, owing to formation of new covalent bonds between C adatom and graphene the isosurface charge density of C-C bonds presented in Fig. \ref{Figure_CF} (a) mimics the isosurface in the case of oxygen adsorbed on graphene as presented in Fig. \ref{Figure1} (d). The nearby bonds of $B_{C1}$ and $B_{C2}$ contain more electronic charge compared to $B_{C3}$ and other C-C bonds as shown in Fig. \ref{Figure_CF} (a). Moreover, it was argued that local disturbances on graphene are long ranged.\cite{longrange,bacsi,lehtinen,can} Interestingly, when a second carbon atom is adsorbed at the close proximity of a second carbon adatom, the bonds of the first one are broken and subsequently it is attached on top of the second adsorbed C atom to form C$_2$ molecule. This way $\sim$ 5 eV energy is gained. The growth of C$_n$ atomic chain continues whenever an adsorbed carbon atom approaches the existing C$_{n-1}$ chain, whereby the chain is detached from graphene and attached to the top of adsorbed carbon atom. These results confirm the earlier study on the perpendicular growth of C$_n$ chains on graphene.\cite{can_chain} While attractive interaction between adsorbed carbon atoms on graphene give rise to the growth of chains on graphene, the repulsive interaction between oxygen adatoms hinders the formation of O$_2$ molecules. \begin{figure} \includegraphics[width=8cm]{Figure_CF.png} \caption{(Color Online) (a) The bonding configuration of a single carbon atom on graphene and the resulting redistribution of bond charges shown by isosurfaces. (b) The bonding configuration of a single fluorine adatom on graphene with energetically favorable top site. The resulting charges of C-C bonds at close proximity are shown by isosurfaces. (c) The growth pattern in the course of the fluorination of graphene.} \label{Figure_CF} \end{figure} The atomic structure and isosurface charge density profile of fluorine atom adsorbed on graphene is presented in Fig. \ref{Figure_CF} (b). For the case of fluorine adsorption, the energetically favorable site is the top site as shown in Fig. \ref{Figure_CF} (b). The binding energy of single F atom adsorbed on a (4x4) graphene is calculated within LDA and was found to be 2.71 eV.\cite{CF2} However, present calculations using PBE+vdW correction\cite{grimme} yield a binding energy of 1.99 eV for adsorption of single F atom on a (5x5) graphene supercell. Upon fluorine adsorption, the bond charge of nearby atoms is modified as shown in isosurface profile. The Bader analysis yields a charge transfer of 0.57 electrons from carbon atoms to the adsorbed fluorine atom at the top site and this value is also significantly close to the value of charge transfer between carbon and oxygen atoms in the present study. Similar to C and O adsorption, the nearby bonds of $B_{F1}$ and $B_{F2}$ contain more electronic charge compared to $B_{F3}$ and other C-C bonds. The nearest top site between $B_{F1}$ and $B_{F2}$ bonds and its other two analogues around F atoms contain more electronic charge and it turns out that these are energetically most favorable sites for adsorption of additional F atoms. In Fig. \ref{Figure_CF} (c) we present how F atoms cover graphene. The second F atom is bonded to the top site formed by $B_{F1}$ and $B_{F2}$ bonds at the other side of graphene, which is most favorable site compared to to others. The third and fourth F atoms are also bound to other two analogues of this site. The energetics of binding structure are in complete agreement with the amount of bond charges of nearby top sites. The final arrangement containing 10 F atoms show a well defined pattern and further fluorination will be continuation of this pattern. \section{ Oxygen - Oxygen Interaction } The interaction between two free oxygen atoms in vacuum is attractive and the formation of an oxygen molecule is energetically more favorable. We set the total energy to zero when the distance $d_{O-O}$ between them is 7 \AA. Figure \ref{Figure_O-O} (a) shows the variation of the energy with the distance, $d_{O-O}$, between two oxygen atoms. The energy does not vary until $d_{O-O}$ is 3.5 \AA, but it starts to decrease as $d_{O-O}$ decreases and passes through a minimum for $d_{O-O}=1.21$ \AA~. This minimum corresponds to the equilibrium bond length of O$_2$ molecule with a binding energy of 6.67 eV. The process is exothermic and occurs without any energy barrier. However, the situation is rather different when one of the oxygen atom is adsorbed to the graphene surface and the other one is free, but approaching from above towards it. In this case, the position of free oxygen is fixed at preset heights while it is approaching, the rest of the system consisting of adsorbed oxygen and all graphene atoms are fully relaxed within conjugate gradient method. We label some of the stages by letters, A-B-C-D-E, while the two oxygens are approaching each other as shown in Fig. \ref{Figure_O-O} (b). The O-O coupling is initially negligible at large $d_{O-O}$ at A, but it passes through a minimum by lowering 0.5 eV at point B corresponding to $d_{O-O} = 2.63$ \AA. Further decrease of $d_{O-O}$ increases the energy increases until the point C, which is $\sim$ 0.5 eV above the point B. Beyond C, oxygen atom flips sideways at D. If one prevents oxygen atom from flipping by fixing its \textit{x-y}-position, but forces it towards the oxygen atom adsorbed on graphene, the adsorbed one is desorbed and two oxygen atoms form O$_2$ molecule at E. In this exothermic process, once the barrier is overcame, the energy decreases by $\sim$ 3.5 eV. \begin{figure} \includegraphics[width=8cm]{Figure_O-O.png} \caption{(Color Online) (a) The interaction energy between two free oxygen atoms approaching each other from a distance. The distance between them is $d_{O-O}$. (b) The interaction energy between a single oxygen atom adsorbed at the bridge site and a free oxygen atom approaching from the top. Different positions of approaching O atom are shown in the side views. (c) Variation of the energy between two oxygen adatoms on graphene. Some of the positions of the approaching oxygen atom on the path of minimum energy barrier are labeled by numerals ( I-VII ). Top and side views of the configuration of two oxygen atoms are shown by insets.} \label{Figure_O-O} \end{figure} For the case of two oxygen adatoms, both adsorbed on graphene and approaching towards each other, the variation of interaction energy is given in Fig. \ref{Figure_O-O} (c). Some of the positions of the approaching oxygen atom on the path of minimum energy barrier are labeled by numerals. For an oxygen starting from a bridge site at I and approaching towards the other oxygen, the energy shows an oscillatory behavior. The maxima, such as II, correspond to the positions where the approaching oxygen is at top site, while the minima, such as I, III, IV, V, VI correspond to positions at the bridge site. The charges of bond charge at the bridge site for reasons discussed before result in the changes in the energies at the bridge sites. For example, the bridge site VI contains more electronic charge and hence it marks the lowest energy position as one oxygen adatom is approaching the other oxygen adatom. The energetics of diffusion through the path between V and VI and energy barrier between them is in good agreement with the calculations by Sun and Fabris.\cite{sun} Since our objective is to investigate the desorption of adsorbed oxygen from graphene surface, we did not consider the energetics of diffusion from site VI to the bridge position on the C-C bond, which is parallel to the C-C bond holding the other oxygen. However, in Ref[\onlinecite{sun}] the barrier to jump to this site is higher. Beyond the point VII, the non-magnetic oxygen/graphene system acquires finite magnetic moments of $\approx$ 0.3 $\mu_B$. Due to the repulsive interaction between two oxygen atoms adsorbed on graphene the energy increases by $\sim$3.2 eV as shown in the Fig. \ref{Figure_O-O} (c). Eventually, the approaching oxygen atom is released from the graphene when the energy barrier is overcame. The final structure is shown by inset. These results indicate that the binding energy of each oxygen on graphene is quite strong and the formation of oxygen molecule as a result of two oxygen atom approaching each other requires significant energy barrier to overcome. \section{ Electronic properties varying with oxygen coverage } The electronic energy structure of GOX strongly depends on oxygen coverage, as well as on the pattern of coverage. Here we consider the electronic properties corresponding to different number of oxygen atoms adsorbed at different bridge sites of the (4x4) supercell repeating periodically. Bare graphene has a semimetallic electronic structure with its characteristic density of states (DOS) making a dip at the Fermi level and linearly crossing valance and conduction bands at special K- and K'-points of Brillouin Zone as shown in Fig. \ref{Figure3} (a). It has a zero band gap and these special symmetry points are called Dirac points.\cite{dirac} The energetically favorable configuration of four oxygen atoms adsorbed on a (4x4) hexagonal supercell is presented in Fig. \ref{Figure3} (b). The resulting DOS profile is different from the bare graphene, since a narrow energy gap of 70 meV is opened. The Dirac cones disappeared and the band gap occurs at the points different than K- and K'-points. For a random and energetically less favorable distribution of oxygen atoms as in Fig. \ref{Figure3} (c), the energy band gap is further increased to 127 meV. The position of the minimum of conduction band and the maximum of the valence band has changed in BZ. Surprisingly, the semimetallic band structure of graphene is recovered when four oxygen atoms are uniformly disturbed on graphene surface as shown in Fig. \ref{Figure3} (d). Although the difference of the atomic positions from Fig. \ref{Figure3} (c) is minute, the band gap is closed and the density of states profile becomes similar to that of bare graphene making a dip at the Fermi level. The conduction and valance bands cross at K- and K'-special points similar to bare graphene. \begin{figure} \includegraphics[width=8cm]{Figure3.png} \caption{(Color Online) (a) Bare graphene and its typical density of states with zero state density at the Fermi level E$_F$. The constant energy surfaces of conduction and valence bands are shown on left-hand side. (b) A four-adatom domain corresponding to lowest total energy configuration and has a band gap of 70 meV. (c) Another adsorption configuration of four oxygen adatom resulting in a band gap of 127 meV. (d) A uniform and symmetric configuration of adsorbed oxygen atoms with zero density of states at $E_F$. (e) A sizable band gap is opened when the symmetry of oxygen decoration is broken by the removal of a single oxygen atom. (f) The wide band gap of 3.25 eV is opened for coverage corresponding to $\Theta = N_{O}/N_{C} = 0.5$ at one side (N$_O$ and N$_C$ are the numbers of oxygen and carbon atoms in the (4x4) supercell.} \label{Figure3} \end{figure} Earlier, it was reported that the superstructures and nanomeshes having special point-group symmetry, which are generated by decoration of adatoms, adatom groups or holes repeating periodically in graphene matrix may give rise to the linearly crossing bands and hence to the recovery of massless Dirac Fermion behavior.\cite{hasan-nanomesh} Our results in Fig. \ref{Figure3} (d) is a verification of this situation for oxygen adsorption on graphene. However, the perfect uniform coverage of oxygen on graphene is experimentally not achievable and we also present a situation where the periodicity of uniform coverage is broken by removal of an oxygen atom as in Fig. \ref{Figure3} (e). In contrast to electronic structure as in Fig. \ref{Figure3} (d), the Dirac behavior is completely removed and the resulting structure is semiconductor with relatively large energy band gap of 501 meV. For the case of $N_{O}/N_{C}$=0.5 coverage at one side, the resulting system is a wide band gap material. Fig. \ref{Figure3} (f) shows the atomic structure and density of states profile. Unlike bare graphene and low oxygen coverage, the resulting structure has a band gap of 3.25 eV. The two sided coverage for $N_{O}/N_{C}$=0.5 also yields similar DOS profile with a band gap wider than 3 eV. These results indicate that the regions of GOX where each carbon atom is bonded with an oxygen should be an insulator and hence should reflect light. \section{ Discussions and Conclusions } Our study dealt with the adsorption of single and multiple oxygen atoms to graphene surface and explored their desorption. We showed that O$_2$ molecule can merely be physisorbed to graphene surface. In contrast, free oxygen atoms are adsorbed at the bridge sites above C-C bonds by forming strong chemical bonds. Significant amount of charge is transferred to oxygen adatom from graphene, which disturbs the charge distribution of the C-C bonds at the proximity of adsorbate. Additional oxygen atoms are adsorbed to the bridge sites above the C-C bonds of graphene, which has highest charge density. This behavior promotes the developments of domains of oxygen adatoms. The domain pattern which, in fact is energetically favorable is also preserved for oxygen atoms adsorbed to both sides of graphene. The binding energy of adsorbed oxygen atoms display an oscillatory change; it starts from 2.43 eV and eventually raises to 2.80 (3.34) eV at one sided (two sided) full coverage with $\Theta$=0.5. Accordingly, the formation energy of adsorbed oxygen is negative. Even if the sequential adsorption of oxygen atoms forms domains with nonuniform coverage, full coverage can form eventually. High oxygen coverage only at one side of graphene causes to severe deformations of graphene lattice. While the adsorption configurations which can be precursors of unzipped of graphene are not favorable for low coverage of large graphene surfaces, they may occur at the edges of domains comprising large number of oxygen atoms, where underlying graphene lattice is severely distorted. Nonequilibrium conditions occurring at finite temperatures and size effects originating from the small size of underlying graphene may favor the nucleation of precursors of unzipping. Single oxygen migrates on a pathway of minimum energy barrier of 0.6 eV over the honeycomb structure between bridge and top sites. For the same reason the interaction between two oxygen adatoms exhibits an oscillatory variation, but becomes increasingly repulsive as the distance decreases beyond a threshold value. This repulsive interaction hinders desorption of oxygen through the formation of O$_2$ molecule despite the negative formation energy of adsorbed oxygen atoms. The electronic structure of oxidized graphene is strongly dependent on the coverage of oxygen and its configuration. While the massless Dirac Fermion behavior with linearly crossing bands at the Fermi level is maintained for specific coverage conserving certain rotation symmetry, the band gap opens and develops with increasing coverage of oxygen adatom. As oxidized domains dominate the surface, semimetallic bare graphene is transformed into a semiconducting material. It appears that the band gap can be engineered through oxygen coverage. Bright and dark spots observed experimentally on GOX surfaces are expected to be related with metallic and light reflecting semiconducting regions, respectively. We believe that metallic regions corresponds to $sp^2$-bonding regions of graphene. Our results indicate that a specific external effect is required for the fast and reversible transition between metallic and semiconducting states of graphene oxide. \section{ Acknowledgements } This work is supported by TUBITAK through Grant No:108T234. All the computational resources have been provided by TUBITAK ULAKBIM, High Performance and Grid Computing Center (TR-Grid e-Infrastructure). S. C. acknowledges the partial support of TUBA, Academy of Science of Turkey.
\section{Introduction} Study of the mechanism of the particle number production is a very important issue in baryogenesis and leptogensis. In experiments, there are many phenomena which violate the particle number. Such phenomena includes $B \bar{B}$ mixing and neutrino flavor oscillation. While they can be treated with time evolution of pure state, the baryogenesis and leptogenesis occur in the environment where the statistical treatment is suitable. This is because they occur when the definite particle number of the universe is unknown. In this study, We use the non-equilibrium field theory with the density matrix and study the time evolution of the particle number. This paper is organized as follows. In section II, we propose a particle number violating model which consists of a heavy neutral scalar and one complex scalar. In the next section, the current associated with the particle number is written in terms of a Green function of non-equilibrium field theory. In section III, the particle number production rate is computed and its property is discussed. The final section is devoted to conclusion. \section{Lagrangian for the scalar model and Particle Number Production} In the previous paper \cite{Hotta:2012hi}, we proposed the following model for particle number production with the interaction. \begin{eqnarray} {\cal L}&=&\frac{1}{2} \partial_\mu N \partial^\mu N -\frac{M_N^2}{2} N^2 +\nonumber \\ &+& \partial_\mu \phi^\dagger \partial^\mu \phi- m_{\phi}^2 \phi^\dagger \phi+A_\phi N \phi^\dagger \phi \nonumber \\ &+& B^2 \phi^2 +A N \phi^2 + h.c., \end{eqnarray} where $N$ is a real scalar and $\phi$ is a complex scalar. There are two types of the interaction. One is particle number conserving interaction which coefficient is given by $A_\phi$ and the other is particle number violating interaction with the coefficient $A$. There are also two types of mass term. One of them with the coefficient $B^2$ violates the particle number and the other one is a particle number conserving one given by the mass term $m_\phi^2 \phi^\dagger \phi$. One may take a phase convention that $B^2$ is real and $A$ is complex. We denote the phase A as $\phi_A={\rm arg.}A$ and it is a source of CP violation. The mass term $B^2$ breaks U(1) symmetry and it splits one complex scalar fields into the two mass eigenstates of real scalars. Introducing two real scalars as $\phi=\frac{1}{\sqrt{2}}(\phi_1+i \phi_2)$, the Lagrangian is rewritten as, \begin{eqnarray} {\cal L}&=&\frac{1}{2}(\partial_\mu N \partial^\mu N -M_N^2 N^2) \nonumber \\ &+&\frac{1}{2}\sum_{i=1}^2(\partial_\mu \phi_i \partial^\mu \phi_i -m_i^2 \phi_i^2)+ \sum_{ij}\phi_i {\cal A}_{ij} \phi_j N, \end{eqnarray} with $m_1^2=m_\phi^2-B^2, m_2^2=m_\phi^2+B^2$. ${\cal A}_{ij}$ is a two by two matrix and is given by, \begin{eqnarray} {\cal A}= \begin{pmatrix} |A| \cos \phi_A+\frac{A_\phi}{2} & -|A| \sin \phi_A \\ -|A| \sin \phi_A & -|A|\cos \phi_A + \frac{A_\phi}{2} \end{pmatrix}. \end{eqnarray} \section{Computing the current for U(1) charge; particle number} The particle number associated with the complex field $\phi$ is a U(1) current. \begin{eqnarray} j_\mu&=&i (\phi^\dagger \partial_\mu \phi -\partial_\mu \phi^\dagger \phi)\nonumber \\ &=& \phi_2 \partial_\mu \phi_1-\partial_\mu \phi_2 \phi_1. \end{eqnarray} We compute the divergence of the U(1) current with some initial condition specified with the density matrix. \begin{eqnarray} \langle j_\mu(X) \rangle ={\rm Tr}\left(j_\mu(X) \rho(0)\right), \label{eq:U1a} \end{eqnarray} where $\rho(0)$ represents an initial quantum statistical state; \begin{eqnarray} \rho(0)=\frac{\exp[-\beta (H_0-\mu L)]}{{\rm Tr}[\exp[-\beta (H_0-\mu L)]]}. \end{eqnarray} $L$ is U(1) charge given by $L=\int d^3 x j_0$ and $\beta$ is $\frac{1}{T}$ with the temperature $T$. $\mu$ is chemical potential. We assume that at $t=0$ all the interaction terms including U(1) breaking mass term are zero and $t>0$, suddenly they are switched on. Therefore at t=0, Hamiltonian $H_0$ and particle number $L$ commute as; \begin{eqnarray} [H_0, L]=0, \label{eq:com} \end{eqnarray} and at later time $t>0$, the U(1) breaking terms are switched on and the particle number is not conserved. \begin{eqnarray} [H, L] \ne 0. \end{eqnarray} In $t=0$, Eq.(\ref{eq:com}), the density matrix can be written by the following product, \begin{eqnarray} \rho(0)=\frac{\exp(-\mu L)\exp(-\beta H_0)}{{\rm Tr}[\exp(-\mu L)\exp(-\beta H_0)]}. \end{eqnarray} With the property Eq.(\ref{eq:com}), one can write the density matrix even for non-zero chemical potential case. Using the density matrix, the initial particle number at $t=0$ is given as, \begin{eqnarray} \langle L(0) \rangle &\equiv& {\rm Tr} L(0) \rho(0) \nonumber \\ &=& \int \frac{ V d^3 k}{(2 \pi)^3} \frac{\sinh \mu \beta} {\cosh \beta \omega_k-\cosh \beta \mu}. \end{eqnarray} When chemical potential is zero, the initial particle number is zero. In the following, we use the density matrix of $\mu=0$ and focus on the particle number production with the interaction. \section{Green function and U(1) current} $U(1)$ current defined in Eq.(\ref{eq:U1a}) can be written in terms of the Green function of non-equilibrium field theory; \cite{Bakshi:1962dv},\cite{Bakshi:1963bn},\cite{Keldysh:1964ud}, \begin{eqnarray} G_{12}^{12}(x,y)&=&< \phi_1(x) \rho(0) \phi_2(y)> \nonumber \\ &=&{\rm Tr} (\phi_2(y) \phi_1(x) \rho(0)). \end{eqnarray} The lower indices distinguish the species of the light scalar fields; $\phi_i (i=1,2)$. The upper indices distinguish whether the operator is on the time ordered path $1$ or on anti-time ordered path $2$ in closed time path formulation \cite{Schwinger:1960qe}. Using the definition, the divergence of the averaged current with the density matrix is the production rate per unit time and unit volume. This can be related to the Green function as follows, \begin{eqnarray} &&\frac{\partial}{\partial X^\mu} \langle J^\mu(X) \rangle=(\Box_x-\Box_y) G^{12}_{12}(x,y)\Bigr{|}_{x=y=X}.\nonumber \\ \end{eqnarray} If the space translation invariance holds, the current depends on only time. If this is the case, the divergence of the current is equal to time derivative of the particle number density, \begin{eqnarray} \frac{\partial}{\partial X^\mu} \langle J^\mu(X) \rangle = \frac{\partial}{\partial X^0} \langle J^0(X^0) \rangle. \end{eqnarray} With the help of 2 PI (Two Particle Irreducible) formulation of the non-equilibrium quantum field theory \cite{Calzetta:1986ey}, we derive Schwinger Dyson equation for the Green functions in \cite{Hotta:2012hi}. The Schwinger Dyson equation can be solved perturbatively and the divergence of U(1) current obtained in one-loop level. We find, \begin{eqnarray} &&\frac{\partial}{\partial X^\mu} \langle J^\mu(X) \rangle \simeq -4|A| \sin \phi_A A_\phi \nonumber \\ && \int \frac{d^3 p}{2 \omega_p (2 \pi)^3} \int \frac{d^3 k}{2 \omega_k (2 \pi)^3} \frac{1}{2 \omega_N} \nonumber \\ &&\{(n_k+1) n_N (n_p+1)-n_k (n_N+1) n_p \} \nonumber \\ && \{I(\omega_{2p}+\omega_{2k}-\omega_N,X^0)\nonumber \\ &&- I(\omega_{1p}+\omega_{1k}-\omega_N,X^0) \}. \label{eq:prd} \end{eqnarray} where we have ignored the terms which is proportional to $B^2$ and $\omega_{i k}=\sqrt{m_i^2+k^2} (i=1,2)$. The time dependent function $I$ is given as, \begin{eqnarray} I(\Omega,X^0)=\frac{\cos \Omega X^0-1}{\Omega}, \end{eqnarray} where $n_N$ ,$n_k$($n_p$) denote the thermal equilibrium distribution function for N and $\phi$ given as, \begin{eqnarray} n_N&=&\frac{1}{\exp(\beta \omega_N(k+p))-1}, \nonumber \\ n_k&=&\frac{1}{\exp(\beta \omega_k)-1}. \end{eqnarray} In the divergence of the current, first we study the factor related to distributions, \begin{eqnarray} (n_k+1) n_N (n_p+1)-n_k (n_N+1) n_p. \end{eqnarray} The first term implies the decay $N$ to two light scalars while the second term implies inverse decay. If the energy conservation in the following sense, holds, \begin{eqnarray} \omega_k+\omega_p=\omega_N(p+k), \label{eq:econv} \end{eqnarray} then, \begin{eqnarray} (n_N+1) n_k n_p =n_N(n_k+1)(n_p+1). \end{eqnarray} The particle number production is cancelled between the decay process and inverse decay process. Therefore the net particle number can not be produced if the energy conservation of Eq.(\ref{eq:econv}) is satisfied. Next we study the coupling constants in Eq.(\ref{eq:prd}). The production rate is proportional to CP violating phase $\phi_A$. In addition to CP violation, both types of the interactions, one is CP conserving and particle number conserving one $A_\phi$ and the other is CP violating and particle number violating one $A$ are required. Finally we study time dependence. We are interested in the time length for which coherence of the two amplitudes in Eq.(\ref{eq:prd}) corresponding to the $N \rightarrow \phi_i(k) \phi_i(p) $ $(i=1,2)$ is not lost yet. When the coherence remains, one can use the approximation, \begin{eqnarray} &&I(\omega_{2p}+\omega_{2k}-\omega_N,X^0)- I(\omega_{1p}+\omega_{1k}-\omega_N,X^0) \nonumber \\ &&\simeq \frac{\sin \Omega_0 X^0}{\Omega_0} \sin \{(\frac{B^2}{2\omega_k}+\frac{B^2}{2 \omega_p}) X^0\}, \end{eqnarray} where $\Omega_0=\omega_N-\omega_k-\omega_p$. We have used the approximation, \begin{eqnarray} \omega_{2k}=\sqrt{k^2+m_\phi^2+B^2}&=&\omega_k+\frac{B^2}{2 \omega_k}, \nonumber \\ \omega_{1k}=\sqrt{k^2+m_\phi^2-B^2}&=&\omega_k-\frac{B^2}{2 \omega_k}. \nonumber \end{eqnarray} We consider when $X^0$ is large and is proportional to $\frac{1}{B^2}$. In the limit of small $B^2$ with fixed $B^2 X^0$, the time dependent factor becomes, \begin{eqnarray} &&I(\omega_{2p}+\omega_{2k}-\omega_N,X^0)- I(\omega_{1p}+\omega_{1k}-\omega_N,X^0) \nonumber \\ &&\simeq \pi \delta(\Omega_0) \sin\{(\frac{B^2}{2\omega_k}+\frac{B^2}{2 \omega_p}) X^0\}. \end{eqnarray} Substituting the time dependent factor, the divergence of the current is, \begin{eqnarray} &&\frac{\partial}{\partial X^\mu} \langle J^\mu(X) \rangle \simeq -4|A| \sin \phi_A A_\phi \nonumber \\ && \int \frac{d^3 p}{2 \omega_p (2 \pi)^3} \int \frac{d^3 k}{2 \omega_k (2 \pi)^3} \frac{1}{2 \omega_N} \pi \delta(\omega_N-\omega_k-\omega_p) \nonumber \\ &&\{(n_k+1) n_N (n_p+1)-n_k (n_N+1) n_p \} \nonumber \\ &&\times \sin\{(\frac{B^2}{2\omega_k}+\frac{B^2}{2 \omega_p}) X^0\}. \end{eqnarray} The presence of the delta function implies the energy conservation $\omega_N=\omega_k+\omega_p$ holds. Therefore the decay contribution and inverse decay contribution are cancelled each other. Below, we consider only the decay contribution. \begin{eqnarray} &&\frac{\partial}{\partial X^\mu} \langle J^\mu(X) \rangle \Bigr{|}_{\rm Decay} \simeq -4 \pi |A| \sin \phi_A A_\phi \nonumber \\ && \int \frac{d^3 p}{2 \omega_p (2 \pi)^3} \int \frac{d^3 k}{2 \omega_k(2 \pi)^3} \frac{1}{2 \omega_N} \delta(\omega_N-\omega_k-\omega_p) \nonumber \\ &&\frac{e^{\beta(\omega_k+\omega_p)}}{(e^{\beta \omega_k}-1) (e^{\beta \omega_k}-1)(e^{\beta \omega_N}-1)} \nonumber \\ && \times \sin\{(\frac{B^2}{2\omega_k}+\frac{B^2}{2 \omega_p}) X^0\} \nonumber \\ &&\simeq -\frac{1}{64 \pi^3} |A| A_\phi T^2 \sin \phi_A \nonumber \\ && \int_{\beta m_N}^\infty \frac{dU}{\sinh \frac{U}{2}} \int_0^{V_{max}(U)} dV \frac{\sin \frac{2 \beta B^2 X^0 U}{U^2-V^2}}{\cosh\frac{U}{2}- \cosh \frac{V}{2}}, \nonumber \\ \end{eqnarray} where \begin{eqnarray} U=\beta(\omega_k+\omega_p), \quad V=\beta(\omega_k-\omega_p), \end{eqnarray} and \begin{eqnarray} V_{max}=\sqrt{(1-\frac{4 m_\phi^2}{m_N^2})(U^2-\beta^2m_N^2)}. \end{eqnarray} We introduce the following dimensionless quantities, \begin{eqnarray} \hat{\beta}=\beta m_{\phi}, \hat{X^0}=m_\phi X^0, \hat{B}=\frac{B}{m_\phi}, \hat{m}_N=\frac{m_N}{m_\phi}. \label{eq:dimless} \end{eqnarray} To study the long time behaviour of the production rate $\hat{X}^0 \sim \frac{1}{\hat{B}^2}$, we define the rescaled time $t$, \begin{eqnarray} t=\frac{B^2 X^0}{m_{\phi} \pi} =\frac{\hat{B}^2 \hat{X}^0}{\pi}. \label{eq:rescaledt} \end{eqnarray} Using Eq.(\ref{eq:rescaledt}), one may write the time dependent part as, \begin{eqnarray} F[t, \hat{\beta}]&=&\frac{1}{\hat{\beta}^2} \int_{\hat{\beta}\hat{m}_N}^\infty \frac{dU}{\sinh \frac{U}{2}} \nonumber \\ &\times &\int_0^{V_{max}(U)} dV \frac{\sin \frac{2 \pi \hat{\beta} U t}{U^2-V^2}}{\cosh\frac{U}{2}-\cosh\frac{V}{2}}. \label{eq:td} \end{eqnarray} In Fig.1, we show the time dependent factor Eq.(\ref{eq:td}) of the production rate as a function of the dimensionless rescaled time $ t$ for three inverse temperature $\hat{\beta}=\frac{1}{10},\frac{1}{20}$,and $\frac{1}{30}$. We choose $\hat{m}_N=20$. We can see the the production rate oscillates and the amplitude decreases. As temperature grows, the production rate becomes larger. \begin{figure*}[th]\center \includegraphics[width=7cm]{Fig.eps} \caption{Time dependence of the production rate of the particle number density. The dashed line corresponds to $\hat{\beta}=\frac{1}{10}$, the thick solid line corresponnds to $\hat{\beta}=\frac{1}{20}$ , and the thin solid line corresponds to $\hat{\beta}=\frac{1}{30}$.} \end{figure*} \section{Conclusion} \begin{itemize} \item{We compute the time variation of the particle number with a scalar model.}\item{In the model, a neutral scalar and one complex scalar are included. U(1) charge related to the complex scalar is the particle number.} \item{The particle number and CP symmetry are violated due to the mass term and the interaction.} \item{Due to U(1) soft-breaking term, one complex scalar is not a mass eigenstate and mass eigenstates are two real scalars with non-degenerate mass.} \item{We have computed the time dependence of the production rate of the particle number density.} \item{The rate oscillates and is damping in their amplitude. The oscillation period is larger for the smaller non-degeneracy.} \item{Because of the damping, the particle number produced during the first half cycle of the oscillation, may remain after the integration of the rate with respect to time.} \item{We have not included the effect of the expanding universe, which effect leads to the decay and inverse decay contribution may not cancel.} \item{As an extension of our work, one can consider the case for $<L(0)>\ne0$ and how the initial particle number is washed out. Also we may apply the method to flavored leptogenesis; $L_e(t), L_\mu(t), L_\tau(t)$.} \end{itemize} \section*{Acknowledgement} This work was supported by Grant-in-Aid for Scientific Research (C) ,Grant Number 22540283 from JSPS. We thank organizers and participants of QFTG2012.
\section{introducci\'on} The theoretical and experimental study of the electromagnetic properties of elementary particles have long represented an interesting research topic in particle physics. Along these lines, the study of the magnetic dipole moment (MDM) and the electric dipole moment (EDM) of fermions has attracted considerable attention in the literature. In particular, it is believed that the study of the EDM may shed light on new sources of CP violation. On the other hand, less attention has been paid to the weak properties of fermions, the weak magnetic dipole moment (WMDM) and the weak electric dipole moment (WEDM), which are the analogues of the fermion electromagnetic properties but are associated with the interaction of a fermion with the neutral weak gauge boson. In the standard model (SM), the EDM and the WEDM arise from the CP-violating phase appearing in the Cabibbo-Kobayashi-Maskawa (CKM) matrix \cite{Kobayashi:1973fv}. Although such a phase is enough to explain the CP violation observed in the $K^0-\bar{K}^0$ system \cite{Christenson:1964fg}, it does not account for the baryogenesis of the universe. However, recent evidences of neutrino oscillations \cite{osci} suggest that these particles have nonzero mass, which opens up the possibility for lepton flavor violation (LFV) and a source of CP violation in the lepton sector. Although the electron and muon electromagnetic properties have been measured with high accuracy, our knowledge of the $\tau$ electromagnetic properties is still beyond an acceptable level, which is due mainly to the fact that the $\tau$ lifetime is too short to allow one to directly measure its interaction with an electromagnetic field. However, indirect bounds on the $\tau$ electromagnetic properties have been obtained via the study of the deviations of the cross sections for $\tau$ production at the CERN LEP. For instance, the current constraints on the $\tau$ electromagnetic properties were obtained through the study of the processes $e^+e^- \to \tau^+\tau^- \gamma$ and $e^+e^- \to e^+e^-\tau^+\tau^-$. Limits on the latter reaction allowed the DELPHI collaboration \cite{Abdallah:2003xd} to place the following bounds: \begin{eqnarray} -0{.}052< &a_\tau& < 0{.}013, \\ -0{.}22< & \text{Re}(d_\tau) &< 0{.}45,\\ -0{.}25 <& \text{Im}(d_\tau) &< 0{.}008,\label{etau} \end{eqnarray} where the EDM is expressed in units of $10^{-16}$ e cm. These results are well beyond the theoretical predictions of the SM: $a_\tau^{\text{SM}}=1177{.}21(5)\times10^{-6}$\cite{asm} and $d_\tau^{\text{SM}}<10^{-34}$ e cm \cite{dsm}. On the other hand, the weak properties of the $\tau$ lepton remain almost unexplored up to date, though the first constraints on them were obtained from the study of the cross section for the process $e^+e^- \to \tau^+\tau^-$ by using a center-of-mass energy near the $Z$ resonance \cite{Acciarri:1998zc,Acton:1992ff}. The current bounds on the weak properties of the $\tau$ lepton, which were obtained by the ALEPH collaboration \cite{Heister:2002ik} using a data sample collected from 1990 to 1995 corresponding to integrated luminosity of 155 pb$^{-1}$, are given by: \begin{eqnarray} \text{Re}(a_\tau^{W}) &<& 1{.}1\times10^{-3},\\ \text{Im}(a_\tau^{W})&< & 2{.}7 \times 10^{-3},\\ \text{Re}(d_\tau^{W}) &<& 0{.}5 \times 10^{-17}\,\,\text{e cm},\\ \text{Im}(d_\tau^{W})&< & 1{.}1 \times 10^{-17}\,\,\text {e cm},\label{wtau} \end{eqnarray} which again are far above the SM predictions $a_\tau^{W}=-(2{.}10+ 0{.}61 i)\times 10^{-6}$ \cite{Bernabeu:1994wh} and $d_\tau^{W}<8\times 10^{-34}$ e cm \cite{Booth:1993af}. However, several SM extensions predict large contributions to these observables that are closer to the experimental bounds. We will explore this possibility in the unparticle physics scenario proposed recently \cite{Georgi:2007ek}. In this framework, the gauge group is $SU(N)$ and there is a hidden sector in the low energy regime. It is conjectured that the theory remains conformal in the infrared (IR) regime, in such a way that there is a continuous mass spectrum. In this sense, the particle concept cannot be defined. Such a hidden sector would interact weakly with the SM via the exchange of heavy states and would manifest itself at an energy scale $\Lambda_{\mathcal U} > 1$ TeV. This scenario can have interesting consequences in both theoretical and phenomenological high energy physics. Despite the inherent complexity of the theoretical framework, the unparticle physics effects can be studied via an effective theory. Indirect bounds on the unparticle parameters have been obtained from LFV decays \cite{Moyotl:2011yv}, the muon anomalous magnetic moment \cite{Hektor:2008xu}, and monophoton production plus missing transverse energy at the LEP \cite{Cheung:2007ap}. More recently, experimental evidence of unparticles has been searched for at the CERN LHC \cite{Ask:2008fh}. In particular, the search for monojets plus missing transverse energy in the 2010 LHC run data has allowed the CMS collaboration to impose strong bounds on the unparticle parameters \cite{Chatrchyan:2011nd}. The physics of the $\tau$ lepton is expected to play an important role in the scientific program of present and future particle colliders \cite{Perl:1991gd,Gentile:1995ue,Pich:1997ym}. Because of its relatively large mass, the $\tau$ lepton can decay hadronically. From this class of processes, high precision measurements of several quantities can be extracted, such as the CKM matrix element $|V_{us}|$ and the mass of the strange quark. Also, as a result of its large variety of decay channels, the study of the $\tau$ lepton represents an interesting tool to search for CP violation, LFV, and other new physics effects. Although the ATLAS \cite{Aad:2011kt} and CMS \cite{Chatrchyan:2011nv} collaborations have already reported their first results for $\tau$ production from $Z$ decays, it is expected that the $B$ factories, BABAR \cite{Aubert:2001tu} and BELLE \cite{:2000cg}, collect large samples of data for $\tau^- \tau^+$ production. Furthermore, since these experiments use a center-of-mass energy around the $\Upsilon(4S)$ mass, they could be useful for the study of the electromagnetic properties of the $\tau$ lepton \cite{Bernabeu:2007rr,Inami:2002ur}. The rest of the work is organized as follows. In Sec. II we present an overview of the effective interactions of a spin-0 unparticle with a fermion pair. Section III is devoted to the analytical results for the fermion weak properties induced by a spin-0 unparticle, whereas the numerical results and discussion for the $\tau$ lepton is presented in Sec. IV, where a brief discussion on the bottom and top weak and electromagnetic properties is also included. The conclusions and outlook are presented in Sec. V. \section{Unparticles Physics overview} A toy model based on a scale invariant sector was already proposed some time ago by Banks and Zaks \cite{Banks:1981nn}, but it was only after the work of Georgi \cite{Georgi:2007ek} that high energy physicists became more interested in this idea. Unparticle physics assumes the existence of a scale invariant hidden sector, known as ${\mathcal B}{\mathcal Z}$ sector, which can interact with the SM fields via the exchange of very heavy particles at a very high energy scale ${\mathcal M}_{\mathcal U}$. Below this energy scale, there are nonrenormalizable couplings between the fields of the ${\mathcal B}{\mathcal Z}$ sector and the SM ones. These couplings can be written generically as $ {\mathcal O}_{SM}{\mathcal O}_{{\mathcal B}{\mathcal Z}}/{\mathcal M}_{\mathcal U}^{d_{SM}+d_{{\mathcal B}{\mathcal Z}}-4}$. The dimension of the associated operators are $d_{{\mathcal B}{\mathcal Z}}$ and $d_{SM}$, respectively. Dimensional transmutation occurs at an energy scale $\Lambda_{\mathcal U}$ due to the renormalizable couplings of the ${\mathcal B}{\mathcal Z}$ sector. Below the scale $\Lambda_{\mathcal U}$, an effective theory can be used to describe the interactions between the fields of the ${\mathcal B}{\mathcal Z}$ sector and the SM fields, which arise from the exchange of unparticle fields. The corresponding effective Lagrangian that respects $SU_L(2)\times U_Y(1)$ gauge invariance can be written as \cite{Georgi:2007ek}: \begin{equation} {\mathcal L}_{\mathcal U}=C_{{\mathcal O}_{\mathcal U}} \frac{\Lambda_{\mathcal U}^{d_{{\mathcal B}{\mathcal Z}}-d_{\mathcal U}}}{{\mathcal M}_{\mathcal U}^{d_{SM}+d_{{\mathcal B}{\mathcal Z}}-4}} {\mathcal O}_{SM}{\mathcal O}_{\mathcal U}, \end{equation} where $C_{{\mathcal O}_{\mathcal U}}$ is a coupling constant and the dimension, $d_{\mathcal U}$, of the unparticle operator, ${\mathcal O}_{\mathcal U}$, can be a fractionary number, though its value is restricted to the interval $1<d_\mathcal U<2$ due to unitarity \cite{Georgi:2007ek,Grinstein:2008qk,Biggio:2008in,Nakayama:2007qu}. As far as the unparticle operators are concerned, their Lorentz structure can be constructed out of the operators ${\mathcal O}_{{\mathcal B}{\mathcal Z}}$ and their transmutation. In general there can be unparticle operators of scalar, ${\mathcal O}_{\mathcal U}$, vector, ${\mathcal O}_{\mathcal U}^\mu$, and tensor, ${\mathcal O}_{\mathcal U}^{\mu\nu}$, type. For simplicity we will only consider spin-0 unparticle operators in our analysis below. The effective Lagrangian describing the scalar and pseudo-scalar interactions of a spin-0 unparticle with a fermion pair is given by: \begin{eqnarray} {\mathcal L}_{{\mathcal U}^\text{spin-0}}&=& \frac{\lambda_{ij}^{S}}{\Lambda_{\mathcal U}^{d_{\mathcal U}-1}} \bar{f}_i f_j {\mathcal O}_{\mathcal U} +\frac{\lambda_{ij}^{P}}{\Lambda_{\mathcal U}^{d_{\mathcal U}-1}} \bar{f}_i \gamma^5 f_j {\mathcal O}_{\mathcal U},\label{lup} \end{eqnarray} where $\lambda_{ij}^{{S,P}}=C_{{\mathcal O}_{\mathcal U}} \Lambda_{\mathcal U}^{d_{{\mathcal B}{\mathcal Z}}}/{\mathcal M}_{\mathcal U}^{d_{SM}+d_{{\mathcal B}{\mathcal Z}}-4}$ stands for the respective coupling constant. Constraints on the coupling constant associated with the $\tau$ lepton have been already obtained from the LFV decay $\tau \to 3 \mu$ \cite{Moyotl:2011yv} and the muon anomalous magnetic moment \cite{Hektor:2008xu}. As far as the unparticle propagators are concerned, they are constructed using scale invariance and the spectral decomposition formula. The propagator for a spin-0 unparticle can be written as \begin{equation} \Delta_F(p^2)= \frac{A_{d_\mathcal U}}{2\sin(d_\mathcal U \pi)} (-p^2-i\epsilon)^ {d_{\mathcal U}-2}, \label{unpro} \end{equation} where $A_{d_\mathcal U}$, which is meant to normalize the spectral density \cite{Cheung:2007ap}, is given by \begin{equation} A_{d_\mathcal U}=\frac{16\pi^2\sqrt{\pi}}{(2\pi)^{2d_\mathcal U}}\frac{\Gamma(d_\mathcal U+\frac{1}{2})}{\Gamma(d_\mathcal U-1)\Gamma(2d_\mathcal U)}. \end{equation} As expected, in the limit of $d_\mathcal U \to 1$, Eq. (\ref{unpro}) becomes the propagator of a massless scalar particle. \section{Electromagnetic and weak properties of the fermions} The electromagnetic and weak properties of fermions can be described through the following interaction Lagrangian: \begin{eqnarray} {\mathcal L}^{\text{spin}-1/2}&=&-\frac{i}{2} \bar{f} \sigma_{\mu\nu} \gamma_5 f(d_f F^{\mu\nu}_\gamma+d_f^{W} F^{\mu\nu}_Z) \nonumber\\ {}&{}&+\frac{e}{4m_f}\bar{f} \sigma_{\mu\nu} f(a_f F^{\mu\nu}_\gamma+a_f^{W} F^{\mu\nu}_Z),\label{lad} \end{eqnarray} where $F^{\mu\nu}_\gamma$ and $F^{\mu\nu}_Z$ are the electromagnetic and weak stress tensors, respectively. The fermion electromagnetic and weak properties arise at the loop level and can be extracted from the matrix element $ie\bar{\mathrm{u}}(p') \Gamma^\mu_V \mathrm{u}(p)$, where $ \Gamma_V^\mu$ is given by: \begin{eqnarray} \Gamma^\mu_V(q^2)&=& F_A(q^2)(\gamma^\mu \gamma_5 q^2-2{m_f} \gamma_5 q^\mu)+F_1(q^2) \gamma^\mu \nonumber\\ {}&{}&+F_2(q^2)i\sigma^{\mu\nu} q_\nu+F_3(q^2)\sigma^{\mu\nu} \gamma_5 q_\nu, \label{vad} \end{eqnarray} with $q=p'-p$ the four-momentum of the gauge boson $V$. The MDM and the EDM are given by $a_f=-2m_f F_2(q^2=0)$ and $d_f=-eF_3(q^2=0)$, whereas the weak properties, $a_f^{W}$ and $d_f^{W}$, are defined by analogue expressions but with the replacement $q^2=m_Z^2$. We now consider the flavor changing interaction given by Eq. (\ref{lup}) to obtain the WMDM and WEDM of the fermion $f$ induced by a spin-0 unparticle. We have calculated the loop amplitudes via Feynman parameters. The results can be written as \begin{eqnarray} a_f^{{W}}(d_\mathcal U)&=& \frac{ A_{d_\mathcal U} }{16 \pi^2\sin{(d_\mathcal U} \pi)} \sum_{i=e,\mu,\tau} \sqrt{r_{i}} \bigg( \frac{m_i^2}{\Lambda_{\mathcal U}^2} \bigg)^{d_\mathcal U-1} \int_0^1dx \int_0^{1-x}dy H(d_\mathcal U,r_{i},x_Z,x,y) \big[ g_V^f F_1(r_{i},x)+2 g_A^f F_2(r_{i},x)\big],\label{aw1} \end{eqnarray} and \begin{eqnarray} d_f^{{W}}(d_\mathcal U) &=&\frac{ - e g_V^f A_{d_\mathcal U}}{16 \pi^2m_f\sin{(d_\mathcal U} \pi)} \sum_{i=e,\mu,\tau} \text{Im}\big( {\lambda_{f i}^P}^* \lambda_{f i}^S \big) \sqrt{r_i} \bigg( \frac{m_i^2}{\Lambda_{\mathcal U}^2} \bigg)^{d_\mathcal U-1} \int_0^1dx \int_0^{1-x}dy (1-x) H(d_\mathcal U,r_{i},x_Z,x,y),\label{dw1} \end{eqnarray} where we introduced the short-hand notation $r_{i}=m_f^2/m_i^2$, $i$ stands for the flavor index of the internal fermion, and $g_{A,V}^f$ are the fermion coupling constants to the $Z$ gauge boson. We also introduced the following dimensionless functions: \begin{eqnarray} F_1(r_{i},x)&=& (x-1) \big(|\lambda_{f i}^S|^2(1+x\sqrt{r_{i}})+|\lambda_{f i}^P|^2(x\sqrt{r_{i}}-1) \big),\label{f1} \\ F_2(r_{i},x)&=& \sqrt{r_{i}}\,\text{Re}\big( {\lambda_{f i}^P}^* \lambda_{f i}^S \big)x(1-x), \label{f2} \\ H(d_\mathcal U,r_{i},x_Z,,x,y)&=&x^{1-d_\mathcal U}\left(r_{i}x_Z (x+y-1)y+(1-x)(1-r_{i}x)\right)^{d_\mathcal U-2},\label{h} \end{eqnarray} with $x_Z=m_Z^2/m_f^2$. As expected, the fermion WEDM only receives contributions from the vector coupling $g_V^f$ and it is proportional to $\text{Im}\big( {\lambda_{f i}^P}^* \lambda_{f i}^S \big)$, which is expected as this property violates CP. As a cross-check for our calculation, from Eqs. (\ref{aw1}) and (\ref{dw1}) we can obtain the fermion electromagnetic properties reported in Ref. \cite{Moyotl:2011yv} after the replacements $x_Z=0$, $g_A^f=0$ and $g_V^f=Q_f$ are done. Here $Q_f$ is the fermion electric charge in units of $e$. In the following section we will concentrate on the numerical evaluation of the electromagnetic and weak properties of the $\tau$ lepton, and also comment briefly on the respective properties of the bottom and top quarks. \section{Numerical analysis and discussion} The analysis of monophoton production plus missing transverse energy, $e^+e^-\to\gamma +X$, at the LEP was used in Ref. \cite{Cheung:2007ap} to impose a bound on the scale $\Lambda_{\mathcal U}$ as a function of $d_{\mathcal U}$. They considered the 95 \%C. L. limit $\sigma(e^+e^-\to\gamma +X)\simeq 0{.}2$ pb obtained at $\sqrt{s}=207$ GeV by the L3 Collaboration. It was found that this limit requires $\Lambda_{\mathcal U}\ge 660$ TeV for $d_\mathcal U=1{.}4$ and $\Lambda_{\mathcal U}\ge 1{.}35$ TeV for $d_\mathcal U= 2$. Stronger limits were obtained by the CMS collaboration using the data for monojet production plus missing transverse energy at the LHC for $\sqrt{s}=7$ TeV and an integrated luminosity of 35 pb$^{-1}$. Such data require $\Lambda_{\mathcal U}\ge 10$ TeV for $d_\mathcal U=1{.}4$ and $\Lambda_{\mathcal U}\ge 1$ TeV for $d_\mathcal U=1{.}7$ \cite{Chatrchyan:2011nd}. In summary, the region $d_\mathcal U< 1{.}4$ is strongly constrained as very large values of $\Lambda_{\mathcal U}$ are required. It is worth mentioning that in obtaining these bounds, the authors of Ref. \cite{Cheung:2007ap,Chatrchyan:2011nd} considered that the unparticle coupling constants have a magnitude of the order of unity. For our analysis we will consider the intervals $1{.}7\leq d_\mathcal U\le 2$ for $\Lambda_{\mathcal U}=1$ TeV and $1{.}4\leq d_\mathcal U\le 2$ for $\Lambda_{\mathcal U}=10$ TeV. To get an estimate of the electromagnetic and weak properties of the $\tau$ lepton, we will also need to make some assumptions concerning the magnitude of the coupling constants involved in the calculation. Based on our previous work \cite{Moyotl:2011yv}, we will consider the following hierarchy for the $\tau$ couplings $\lambda_{\tau e}^{S,P}<\lambda_{\tau \mu}^{S,P}\ll\lambda_{\tau \tau}^{S,P}$. It means that we will assume that LFV interactions occur mainly between the $\mu$ and $\tau$ leptons. Therefore, we will neglect the contributions from the $\lambda_{\tau e}^{S,P}$ coupling. In addition, for the flavor conserving couplings we will assume that $\lambda_{\mu \mu}^{S,P}\simeq\lambda_{\tau \tau}^{S,P}$. As far as the numerical values of the coupling constants are concerned, we will consider values that are consistent with the bounds obtained in Ref. \cite{Moyotl:2011yv} from the experimental limits on the muon MDM and the LFV decay $\tau\to3\mu$. In particular we will consider the value $\lambda_{\tau \tau}^{S,P}\simeq 1{.}6$ (1.0), which is consistent with $d_\mathcal U=1{.}7$ (1.4), and $\Lambda_{\mathcal U}=1$ TeV (10 TeV). Also, when analyzing the MDM and WMDN we will consider the following scenarios: \begin{itemize} \item Lone contribution from the scalar coupling: $\lambda_{\tau i}^P=0$ and $\lambda_{\tau i}^S\ne0$. \item Lone contribution from the pseudo-scalar coupling: $\lambda_{\tau i}^P\ne0$ and $\lambda_{\tau i}^S=0$. \item Contribution from both scalar and pseudo-scalar couplings: $\lambda_{\tau i}^S\simeq \lambda_{\tau i}^P$. \end{itemize} In the case of the EDM and WEDM, since they require the simultaneous contribution from both scalar and pseudo-scalar couplings, we will only consider the last scenario. \subsection{$\tau$ magnetic dipole moment} Numerical evaluation of the $\tau$ MDM induced by a spin-0 unparticle shows that the pseudo-scalar contribution is negative whereas the scalar contribution is positive, with the latter slightly larger in magnitude than the former. We have plotted in Fig. \ref{aem1-10} the scalar contribution, the absolute value of the pseudo-scalar contribution, and the total contribution for $\Lambda_{\mathcal U}=1$ TeV and $\Lambda_{\mathcal U}=10$ TeV. We also included the SM prediction, which is given by the horizontal line. Since the unparticle propagator contains the term $\sin(d_\mathcal U\pi)$ in the denominator, the contributions to the MDM diverge when $d_\mathcal U \to 2$. Therefore, in the allowed area, the largest contributions to the $\tau$ MDM are reached for $d_\mathcal U$ around 2. In this case $a_\tau^{\mathcal U}$ can be of the order of the SM contribution or larger. On the other hand, for values of $d_\mathcal U$ close to the lower bound, $a_\tau^{\mathcal U}$ is of the order of $10^{-6}$. When $\Lambda_{\mathcal U}=10$ TeV, the unparticle contribution to the $\tau$ MDM is more suppressed and its lowest values are of the order of $10^{-9}-10^{-10}$ . Since the scalar and pseudo-scalar contributions to $a^{\mathcal U}_\tau$ are about the same order of magnitude but opposite in sign, the total contribution can cancel out largely, which is more evident for $d_\mathcal U$ around 1.9. Therefore the largest contribution to $a^{\mathcal U}_\tau$ would arise in the scenario when only one contribution, scalar or pseudoscalar, is present and for low values of $\Lambda_\mathcal U$. For $d_\mathcal U$ around $1{.}95$, both scalar and pseudo-scalar contributions reach their minimal absolute values: $|a_\tau^\mathcal U|\simeq 5\times10^{-7}$ when $\Lambda_{\mathcal U}=1$ TeV and $|a_\tau^\mathcal U|\simeq3\times10^{-9}$ when $\Lambda_{\mathcal U}=10$ TeV. \begin{figure}[!hbt] \centering \includegraphics[width=8.5cm]{aem1.eps} \includegraphics[width=8.5cm]{aem10.eps} \caption{Contribution from a spin-0 unparticle to the $\tau$ MDM as a function of the scale dimension $d_\mathcal U$ for two values of $\Lambda_\mathcal U$. We show the pure scalar contribution (solid line), the absolute value of the pure pseudo-scalar contribution (dashed line), and the total contribution (dotted line). The horizontal line is the SM contribution, and the vertical line represents the lower bound obtained by the CMS collaboration \cite{Chatrchyan:2011nd}.} \label{aem1-10} \end{figure} It is interesting to make a comparison between our results and those arising in other SM extensions, such as the SeeSaw model and an extension of the MSSM with a mirror fourth generation. While the type-I and type-III SeeSaw models predict the contributions $|a_\tau^{I}|<1{.}87\times10^{-8}$ and $|a_\tau^{III}|<7{.}55\times10^{-9}$ \cite{Biggio:2008in}, for representative values of the model parameters, the extension of the MSSM with a mirror fourth generation predicts a positive contribution, of the order of $10^{-6}-10^{-9}$ \cite{Ibrahim:2008gg}. The $\tau$ MDM has also been studied in the framework of the effective Lagrangian approach and the Fritzsch-Xing lepton mass matrix, but the respective contributions are even more suppressed \cite{adtau-eff}, of the order of $10^{-11}$. We thus conclude that the contribution from a spin-0 unparticle to the $\tau$ MDM can be of the same order of magnitude and even larger than the predictions of other SM extensions. \begin{figure}[!hbt] \centering \includegraphics[width=8.5cm]{dem1.eps} \includegraphics[width=8.5cm]{dem10.eps} \caption{Absolute values of the real (solid line) and imaginary (dotted line) parts of the contribution from a spin-0 unparticle to the $\tau$ EDM as a function of the scale dimension $d_\mathcal U$ for two values of $\Lambda_\mathcal U$. The horizontal line is the SM contribution, and the vertical line represents the lower bound obtained by the CMS collaboration \cite{Chatrchyan:2011nd}.} \label{dem1-10} \end{figure} \subsection{$\tau$ electric dipole moment} The EDM requires an internal fermion in the loop different than the external fermion, so it must be induced by flavor changing couplings. Furthermore, it is necessary the presence of a CP-violating phase in the constant couplings. We thus write $\text{Im}\big( \lambda_{\tau\mu}^{P*} \lambda_{\tau\mu}^S \big)=|\lambda_{\tau\mu}^S||\lambda_{\tau\mu}^P| \sin\Delta \Phi_{\tau\mu}^{S,P}$, where $\Delta\Phi_{\tau\mu}^{S,P}=\theta_{\tau\mu}^S-\theta_{\tau\mu}^P$ is the relative phase between the scalar and pseudo-scalar couplings. It is only the relative CP-violating phase that must be nonzero in order to have an EDM. Depending on this phase, the EDM can be negative or positive, which poses no problem as the experimental bound also comprehends negative values. In order to analyze the unparticle contribution to the $\tau$ EDM, we will not consider specific values for $\text{Im}\big( \lambda_{\tau\mu}^{P*} \lambda_{\tau\mu}^S \big)$. In Fig. \ref{dem1-10} we have plotted the absolute values of the real and imaginary parts of the contribution to the $\tau$ EDM from a spin-0 unparticle as a function of the scale $d_\mathcal U$ for $\Lambda_{\mathcal U}=1$ TeV and $\Lambda_{\mathcal U}=10$ TeV. A detailed analysis allows us to conclude that there is a change in the sign of the real part of the $\tau$ EDM at $d_\mathcal U\simeq 1{.}325$, whereas its imaginary part is always positive. In the allowed region, both the real and imaginary parts are positive, although the CP-violating phase can give an additional change of sign. It is also interesting that the real part diverges when $d_\mathcal U \to 2$, but the imaginary part is negligibly small. Therefore, around $d_\mathcal U=2$, the $\tau$ EDM is almost real and also reaches its largest size. At the lowest allowed value of $d_\mathcal U$, the contributions to the $\tau$ EDM are $d_\tau^{\mathcal U}=\text{Im}\big( \lambda_{\tau\mu}^{P*} \lambda_{\tau\mu}^S \big) (2{.}14+1{.}25 i) \times10^{-21}$ e cm when $\Lambda_\mathcal U=1$ TeV and $d_\tau^{\mathcal U}=\text{Im}\big( \lambda_{\tau\mu}^{P*} \lambda_{\tau\mu}^S \big) (0{.}69+3{.}17 i) \times10^{-20}$ e cm when $\Lambda_{\mathcal U}=10$ TeV. It can also be observed that, in the allowed region, the real part reaches its minimal value at $d_\mathcal U\simeq 1{.}9$, where $d_\tau^{\mathcal U}=\text{Im}\big( \lambda_{\tau\mu}^{P*} \lambda_{\tau\mu}^S \big) (3{.}92+0{.}39 i) \times10^{-22}$ e cm when $\Lambda_{\mathcal U}=1$ TeV, and $d_\tau^{\mathcal U}=\text{Im}\big( \lambda_{\tau\mu}^{P*} \lambda_{\tau\mu}^S \big) (5{.}85+0{.}40 i) \times10^{-24}$ e cm when $\Lambda_{\mathcal U}=10$ TeV. In general, the spin-0 unparticle contribution to $d_\tau$ can be above the SM prediction \cite{dsm} as long as $\text{Im}\big( \lambda_{\tau\mu}^{P*} \lambda_{\tau\mu}^S \big)$ is not too small. As far as other SM extensions are concerned, in an extension of the MSSM with vectorlike multiplets, the contributions to the $\tau$ EDM arise at the one loop level from loops carrying $W$ gauge bosons, charginos (${\tilde{\chi}}_i^\pm$) or neutralinos (${\tilde{\chi}}_i^0$). Since these particles are heavier than the $\tau$ lepton, their contributions to the EDM are purely real and have values ranging from $d_\tau\simeq 6{.}5\times10^{-18}$ e cm to $d_\tau\simeq 3{.}0\times10^{-23} $ e cm \cite{Ibrahim:2010va}. In contrast, the unparticle contribution $d_\tau^{\mathcal U}$ is almost real at $d_\mathcal U\simeq 2$, where it can reach values of the order of $10^{-18}$ e cm, though it tends to be smaller for other $d_\mathcal U$ values. The $\tau$ EDM has also been studied in other SM extensions, but the respective predictions were found to be very small. This is the case of the framework of the Fritzsch-Xing lepton mass matrix, in which $|d_\tau|<2{.}2\times 10^{-25}$ e cm \cite{adtau-eff}. \begin{figure}[!hbt] \centering \includegraphics[width=8.5cm]{aw1.eps} \includegraphics[width=8.5cm]{aw10.eps} \caption{Absolute values of the real and imaginary parts of the contribution from scalar (solid and dotted lines, respectively) and pseudo-scalar (dashed and dot-dashed lines, respectively) unparticle couplings to the $\tau$ WMDM, as a function of the scale dimension $d_\mathcal U$ and for two values of $\Lambda_\mathcal U$. The range of the horizontal axis corresponds to the region still allowed according to CMS \cite{Chatrchyan:2011nd}.} \label{aw1-10} \end{figure} \subsection{$\tau$ weak magnetic dipole moment} Contrary to the case of the MDM, the WMDM does depend on the relative phase $\Delta\Phi_{\tau i}^{S,P}$, though a CP violating phase is not required. Furthermore, since such a phase only appears when there is flavor changing couplings, which can be neglected as compared to the flavor conserving ones, for simplicity we will consider a vanishing $\Delta\Phi_{\tau\tau}^{S,P}$. We will first examine the individual behavior of the scalar and pseudo-scalar contributions. To this end, we show in Fig. \ref{aw1-10} the absolute values of the real and imaginary parts of both the scalar and pseudo-scalar contributions to the $\tau$ WMDM for $\Lambda_{\mathcal U}=1$ TeV and $\Lambda_{\mathcal U}=10$ TeV. In this Figure we can observe that both contributions show a similar behavior although the magnitude of the scalar contribution is slightly larger. Moreover, the largest values of both contributions can arise around $d_\mathcal U \to 2$, similar to what is observed in the MDM. It is also interesting to note that for $\Lambda_{\mathcal U}=1$ TeV, the scalar contributions are negative in the whole $d_\mathcal U$ interval, while the pseudo-scalar contributions are positive, contrary to the case of the MDM. For the same value of $\Lambda_\mathcal U$, we also observe that the magnitude of the $\tau$ WDM can be of the order of $10^{-9}$ in the allowed region of $d_\mathcal U$. The situation changes when $\Lambda_{\mathcal U}=10$ TeV, in which case the real part of the scalar contribution changes sign from positive to negative at $d_\mathcal U \simeq 1{.}49$, whereas the imaginary part remains positive. The real and imaginary parts of the pseudo-scalar contribution also show a similar behavior but they are of opposite sign to their scalar analogues. Thus, when $\Lambda_\mathcal U=10$ TeV, the $\tau$ WDM is purely imaginary at $d_\mathcal U\simeq 1.49$, with a magnitude of the order of $10^{-9}$, whereas for other $d_\mathcal U$ values, the magnitude of the real and imaginary parts fall in the range $10^{-8}-10^{-11}$. \begin{figure}[!hbt] \centering \includegraphics[width=8.5cm]{awphi1.eps} \includegraphics[width=8.5cm]{awphi10.eps} \caption{Absolute values of the real (solid line) and imaginary (dotted line) parts of the total contribution of a spin-0 unparticle to the $\tau$ WMDM as a function of the scale $d_\mathcal U$ for two values of $\Lambda_\mathcal U$. For simplicity we neglect LFV and use $\Delta\Phi_{\tau\tau}^{S,P}=0$. The range of the horizontal axis corresponds to the region still allowed according to CMS \cite{Chatrchyan:2011nd}.} \label{awphi} \end{figure} As mentioned above, the LFV contributions to the $\tau$ WMDM are expected to be subdominant. Since the relative phases of the flavor conserving couplings are zero, we have $\text{Re}\big( {\lambda_{\tau \tau}^P}^* \lambda_{\tau \tau}^S \big)=|\lambda_{\tau \tau}^P| |\lambda_{\tau \tau}^S|$. Therefore, apart from the individual contributions from scalar and pseudo-scalar couplings, there are an interference term that contributes to the WMDM. In Fig. \ref{awphi} we show the absolute values of the real and imaginary parts of the total contribution of a spin-0 unparticle to the $\tau$ WMDM for $\Lambda_{\mathcal U}=1$ TeV and $\Lambda_{\mathcal U}=10$ TeV. We note that while the real part diverges when $d_\mathcal U \to 2$, the imaginary part almost vanishes, which is similar to the behavior of the EDM, as shown in Fig. \ref{dem1-10}. It is also observed that when $\Lambda_{\mathcal U}=1$ TeV, both the real and imaginary parts are positive, but when $\Lambda_{\mathcal U}=10$ TeV only the imaginary part is positive whereas the real part changes from negative to positive at $d_\mathcal U\simeq 1.49$. At the lowest allowed value of $d_\mathcal U$, $a_\tau^{W}\simeq (1{.}15+i1{.}56)\times10^{-8}$ when $\Lambda_\mathcal U=1$ TeV and $a_\tau^{W}\simeq (-0{.}70+i 2{.}59\times)10^{-8}$ when $\Lambda_\mathcal U=10$ TeV. The minimal values of the real contribution to $a_\tau^\mathcal U$ are reached at $d_\mathcal U$ around $1.9$, and they correspond to $a_\tau^{W}\simeq (5{.}61+i1{.}82)\times10^{-9}$ when $\Lambda_\mathcal U=1$ TeV and $a_\tau^{W}\simeq (3{.}19+i 0{.}71\times)10^{-11}$ when $\Lambda_\mathcal U=10$ TeV. Finally, we would like to compare our predictions with the SM prediction \cite{Bernabeu:1994wh}. In the most promising scenario, the unparticle contributions are about two orders of magnitude smaller than the SM contribution, but in a more conservative scenario they are about five orders of magnitude below. The unparticle scenario however allows for both positive or negative contributions. \subsection{$\tau$ weak electric dipole moment} As was the case with the EDM, a nonzero WEDM requires a CP-violating phase, which appears when LFV couplings are present. Thus, in order to analyze this property we will follow the same approach used to calculate the EDM. However, before the numerical evaluation, the analysis of Eq. (\ref{h}) suggests that the $\tau$ WEDM is expected to be smaller than the EDM due to the term proportional to $x_Z$. We show in Fig. \ref{dwt} the behavior of the real and imaginary parts of the $\tau$ WEDM induced by a spin-0 unparticle as a function of $d_\mathcal U$ and for two values of $\Lambda_\mathcal U$. As was anticipated, this property shows a behavior similar to that of the $\tau$ EDM, though it has a smaller magnitude and opposite sign. When $\Lambda_{\mathcal U}=1$ TeV, both the real and imaginary contributions are negative, but when $\Lambda_{\mathcal U}=10$ TeV, the imaginary part is negative whereas the real part changes from positive to negative at $d_\mathcal U \simeq 1{.}5$, where the $\tau$ WEDM is almost imaginary, i.e. $d_\tau^{W}\simeq-i\text{Im}\big( \lambda_{\tau\mu}^{P*} \lambda_{\tau\mu}^S \big)2{.}2\times10^{-24}$ e cm. Contrary to behavior of the EDM, both the real and imaginary parts of the WEDM diverge when $d_\mathcal U\to 2$. Also, at the lowest allowed value of $d_\mathcal U$, $d_\tau^{W}\simeq -\text{Im}\big( \lambda_{\tau\mu}^{P*} \lambda_{\tau\mu}^S \big)(3{.}28 + i4{.}51)\times10^{-24}$ e cm when $\Lambda_{\mathcal U}=1$ TeV and $d_\tau^{W}\simeq\text{Im}\big( \lambda_{\tau\mu}^{P*} \lambda_{\tau\mu}^S \big)(2{.}67-i8{.}24)\times10^{-24}$ e cm when $\Lambda_{\mathcal U}=10$ TeV. When $\Lambda_{\mathcal U}=1$ TeV, the minimal value of the real part is reached at $d_\mathcal U \simeq 1{.}76$, where $d_\tau^{W}=-\text{Im}\big( \lambda_{\tau\mu}^{P*} \lambda_{\tau\mu}^S \big)(3{.}21 + i3{.}01)\times10^{-24}$ e cm, while the minimal value of the imaginary part is reached at $d_\mathcal U \simeq 1{.}9$, where $d_\tau^{W}= -\text{Im}\big( \lambda_{\tau\mu}^{P*} \lambda_{\tau\mu}^S \big)(5{.}21 + i1{.}69)\times10^{-24}$ e cm. On the other hand, when $\Lambda_{\mathcal U}=10$ TeV the minimal values of the real and imaginary parts are reached at $d_\mathcal U \simeq 1{.}86$ and $d_\mathcal U = 1{.}93$, respectively. These minimal values correspond to $d_\tau^{W}= -\text{Im}\big( \lambda_{\tau\mu}^{P*} \lambda_{\tau\mu}^S \big)(7{.}40 + i3{.}48)\times10^{-26}$ e cm and $d_\tau^{W}=-\text{Im}\big( \lambda_{\tau\mu}^{P*} \lambda_{\tau\mu}^S \big)(1{.}09 + i0{.}24)\times10^{-25}$ e cm, respectively. \begin{figure}[!hbt] \centering \includegraphics[width=8.5cm]{dw1.eps} \includegraphics[width=8.5cm]{dw10.eps} \caption{The same as in Fig \ref{dem1-10}, but for the $\tau$ WEDM.} \label{dwt} \end{figure} Although the contribution from a spin-0 unparticle to the $\tau$ WEDM can be up to ten orders of magnitude larger than the SM contribution in the best scenario, there is a difference of about seven orders of magnitude with respect to the experimental limits, although the actual value of $\text{Im}\big( \lambda_{\tau\mu}^{P*} \lambda_{\tau\mu}^S \big)$ can reduce additionally the size of the unparticle contribution or change its sign. In summary, the unparticle scenario allows for negative or positive contributions to the $\tau$ WEDM, which can be closer to the experimental limits than the SM contribution. \subsection{Electromagnetic and weak properties of heavy quarks } Due to quark confinement, the study of the quark properties requires more elaborated experimental techniques. Indirect measurements can be extracted from composite states such as the neutron, the proton, the deuteron, or atoms of thallium and ${}^{199}$Hg. For instance, the MDM of the charm and bottom quarks can be extracted from the heavy baryons $\Sigma_c$, $\Lambda_c$, $\Xi_c$ and $\Xi_b$\cite{dm-hb}. On the other hand, it has been proposed that the top quark properties could be measured at hadron colliders via $ V t\bar{t}$ production, with the top quarks decaying in the dominant channel $t\to W b$. Along these lines, the authors of Ref. \cite{Baur:2004uw} have shown that the LHC would allow one to measure anomalous contributions to the $\gamma t \bar{t}$ coupling via the process $pp\to \gamma \bar{t} t\to \gamma W^+ \bar{b} W^- b$ as long as one of the $W$ gauge bosons decays leptonically $W\to l \nu_l$ and the other one decays hadronically $W\to jj$. When $V=Z$, due to the trigger efficiency, it is assumed that the $Z$ gauge boson decays leptonically $Z\to \bar{l'} l'$, with either one of the $W$ gauge boson decaying leptonically and the other one decaying hadronically or with both of them decaying hadronically. In this scenario, it was shown that the measurements at the LHC would be sensitive enough to allow one to extract anomalous contributions to the $Z t \bar{t}$ couplings from the data on the processes $p p \to l'\bar{l'} l \nu_l b \bar{b} jj$ and $p p \to l'\bar{l'} b\bar{b}+4j$, as long as the luminosity is increased by a 10 factor (SLHC). A more detailed discussion on the technical details of this analysis is beyond the present work, so we refer the reader to the original references. We will content ourselves with analyzing the potential unparticle effects on the electromagnetic and weak properties of heavy quarks, namely the bottom quark and the top quark. We will consider the scenarios discussed above for the unparticle parameters and the coupling constants. Since the behavior of the electromagnetic and weak properties of heavy quarks is similar to the one observed in the $\tau$ lepton case, we will only present an estimate of the unparticle contributions at the lowest allowed value of $d_\mathcal U$. Also, we will present the minimal values of the real part, which are obtained for $d_\mathcal U$ around 1.9. We show the results in Tables \ref{ew-b} for the bottom quark and Table \ref{ew-t} for the top quark. \begin{table}[!htb] \begin{center} \begin{tabular}{|c|cccc|} \hline {} & $\Lambda_{\mathcal U}=1$ TeV, ${d_\mathcal U = 1{.}7}$ & $\Lambda_{\mathcal U}=10$ TeV, ${d_\mathcal U = 1{.}4}$ & $\Lambda_{\mathcal U}=1$ TeV &$\Lambda_{\mathcal U}=10$ TeV \\ \hline \hline $a_b$ & $2{.}55\times10^{-8}$ & $1{.}17\times10^{-7}$ & $2{.}80\times10^{-9}$& $1{.}40\times{10}^{-11}$\\ $d_b/\text{Im}({\lambda_{bs}^P}^*\lambda_{bs}^S)$ & $(1{.}63+i1{.}55)\times10^{-23}$ & $(-0{.}11+i1{.}07)\times10^{-22}$& $(2{.}99+i0{.}39)\times10^{-24}$ & $(1{.}69+i0{.}22)\times10^{-26}$ \\ $a_b^{W}$& $(0{.}73+i1{.}00)\times10^{-8}$ & $(-0{.}42+i1{.}66)\times10^{-8}$& $(3{.}59+i1{.}16)\times10^{-9}$ & $(2{.}05+i0{.}49)\times10^{-11}$\\ $d_b^{W}/\text{Im}({\lambda_{bs}^P}^*\lambda_{bs}^S)$& $(-4{.}34-i5{.}98)\times10^{-26}$ & $(1{.}37-i4{.}23)\times10^{-26}$& $(-6{.}91-i2{.}24)\times10^{-26}$& $(-6{.}69-i1{.}27)\times10^{-28}$\\ \hline \end{tabular} \caption{Contributions from a spin-0 unparticle to the electromagnetic and weak properties of the bottom quark. The values shown are those obtained at the lowest allowed value of $d{\mathcal{U}}$ (second and third columns) together with the values that correspond to the minimal value of the real part (fourth and fifth columns) for two values of $\Lambda_{\mathcal{U}}$.} \label{ew-b} \end{center} \end{table} Unfortunately, there are no other theoretical predictions in the case of the bottom quark to our knowledge. However, due to its heavy mass, it has been suggested that the top quark may be sensible to new physics effects. This has motivated the study of the top quark properties, such as the EDM \cite{Ibrahim:2010hv}, WEDM \cite{Hollik:1998vz}, the chromoelectric dipole moment \cite{Ibrahim:2011im}, and the chromomagnetic dipole moment \cite{Martinez:2008hm}. Despite its heaviness, the top EDM is predicted to be negligibly small in the SM, i.e. $d_t<10^{-30}$ e cm \cite{Soni:1992tn}, which is of the same order of magnitude than that of the $\tau$ lepton. As far as other extensions of the SM are concerned, an MSSM extension with vectorlike multiplets predicts values for $d_t$ ranging from $2{.}87\times10^{-19}$ e cm to $2{.}85\times10^{-22}$ e cm \cite{Ibrahim:2010hv}. The unparticle contribution $d_t^{\mathcal U}$ has an imaginary part due to the internal charm quark but its order of magnitude is below the $10^{-19}$ level, although at best it can be of the same order of magnitude than the contributions predicted in the MSSM extension with vectorlike multiplets. As for the top WEDM, it was calculated long ago in the framework of the R-parity preserving MSSM version with complex parameters, where it was found that $d_t^{W}\simeq(0{.}351-1{.}264)\times10^{-19}$ \cite{Hollik:1998vz}. However, the order of magnitude of this prediction can have a significant decrease if updated bound on the model parameters are considered. On the other hand, the unparticle contribution is much smaller and can be up to five orders of magnitude below. \begin{table}[!htb] \begin{center} \begin{tabular}{|c|cccc|} \hline {} & $\Lambda_{\mathcal U}=1$ TeV, ${d_\mathcal U = 1{.}7}$ & $\Lambda_{\mathcal U}=10$ TeV, ${d_\mathcal U = 1{.}4}$ & $\Lambda_{\mathcal U}=1$ TeV &$\Lambda_{\mathcal U}=10$ TeV \\ \hline $a_t$ & $-2{.}05\times10^{-5}$ & $-7{.}21\times10^{-6}$ & $-1{.}32\times10^{-5}$& $-8{.}52\times{10}^{-8}$\\ $d_t/\text{Im}({\lambda_{tc}^P}^*\lambda_{tc}^S)$ & $(-2{.}42-i2{.}58)\times10^{-20}$ & $(0{.}23-i1{.}32)\times10^{-20}$ & $(-2{.}10-i0{.}72)\times10^{-20}$ & $(-1{.}47-i0{.}32)\times10^{-22}$ \\ $a_t^{W}$ & $-4{.}73\times10^{-5}$ & $-2{.}06\times10^{-5}$& $-2{.}77\times10^{-5}$ & $-1{.}74\times10^{-7}$\\ $d_t^{W}/\text{Im}({\lambda_{tc}^P}^*\lambda_{tc}^S)$& $(4{.}50+i6{.}19)\times10^{-22}$ &$(-0{.}63+i1{.}94)\times10^{-22}$ &$(9{.}34+i3{.}03)\times10^{-22}$& $(6{.}88+i1{.}76)\times10^{-24}$\\ \hline \end{tabular} \caption{The same as in Table \ref{ew-b}, but for the top quark. } \label{ew-t} \end{center} \end{table} \section{Conclusions} We have studied the contribution to the electromagnetic and weak properties of fermions from a spin-0 unparticle, with particular emphasis on the $\tau$ lepton properties. As far as the unparticle parameters are concerned, we considered the most recent CMS bounds from monojet production plus missing transverse energy at the LHC, while for the coupling constants we used the indirect limits obtained from the experimental bounds on LFV decays and the muon MDM. In the most promising scenario, we find that the unparticle contribution to the electromagnetic properties of the $\tau$ lepton can be larger than the contributions predicted by the SM and some of its extensions, such as the SeeSaw model and extensions of the minimal supersymmetric standard model with a mirror fourth generation and vectorlike multiplets. As far the $\tau$ weak properties are concerned, the contributions from a spin-0 unparticle are smaller than the respective contributions to the electromagnetic properties, although they are larger than the SM contributions, tough much smaller than the current experimental limits. We also examine the electromagnetic and weak properties of the bottom and top quarks. In particular, we find that the predictions obtained for the top EDM in the framework of unparticle physics are of similar order of magnitude than in an MSSM extension with vectorlike multiplets. In general, the most promising scenario for the contribution of unparticle physics to the electromagnetic and weak properties of fermions is that in which $d_\mathcal U$ is close to 2, which is a region still allowed by the most recent constraints on unparticle physics from the LHC data. We would like to emphasize however that our results depend considerably on the values of the scale $\Lambda_\mathcal U$ and the dimension $d_\mathcal U$. \acknowledgments{We acknowledge financial support from Conacyt and SNI (M\'exico). G.T.V also acknowledges VIEP-BUAP for partial support.}
\section{The Symbolic Generic Initial System of 6, 7, and 8 Uniform Fat Points in General Position} \label{sec:678} As before, $I \subseteq R[x,y,z]$ will denote the ideal of points $p_1, \dots, p_r$ of $\mathbb{P}^2$ in general position. The goal of this section is to prove the second part of Theorem~\ref{thm:mainthm}. \newtheorem*{thm:mainthmb}{Theorem \ref{thm:mainthm} (b)} \begin{thm:mainthmb} \label{thm:polytopefor678} Suppose that $I \subseteq K[x,y,z]$ is the ideal of $r=6, 7, \text{or } 8$ points of $\mathbb{P}^2$ in general position and that $P \subseteq \mathbb{R}^2$ is the limiting shape of the reverse lexicographic symbolic generic initial system $\{ \text{gin} (I^{(m)})\}_m$. Then the boundary of $P$ is defined by the line segment through the points $(\gamma_1, 0)$ and $(0, \gamma_2)$ where \begin{itemize} \item[(a)] $\gamma_1 = \frac{12}{5}$ and $\gamma_2 = \frac{5}{2}$ when $r=6$; \item[(b)] $\gamma_1 = \frac{21}{8}$ and $\gamma_2 = \frac{8}{3}$ when $r=7$; and \item[(c)] $\gamma_1 = \frac{48}{17}$ and $\gamma_2 = \frac{17}{6}$ when $r=8$. \end{itemize} \end{thm:mainthmb} The proof of this result relies on knowing certain values of the Hilbert functions $H_{I^{(m)}}(t)$ of the ideals $I^{(m)}$ where $I$ is the ideal of 6, 7, or 8 general points. Techniques for computing $H_{I^{(m)}}(t)$ in these cases are not new (for example, see \cite{Nagata60}), but they can be complicated. Thus, we take time in Section \ref{sec:678Background} to review a modern technique for finding the Hilbert functions, and then apply these results to the proof of Theorem \ref{thm:mainthm}(b) in Section \ref{sec:678Proof}. \subsection{Background on Surfaces} \label{sec:678Background} The method we use to compute $H_{I^{(m)}}(t)$ follows the work of Fichett, Harbourne, and Holay in \cite{FHH01}. Suppose that $\pi: X \rightarrow \mathbb{P}^2$ is the blow-up of distinct points $p_1, \dots, p_r$ of $\mathbb{P}^2$. Let $E_i = \pi^{-1}(p_i)$ for $i = 1, \dots, r$ and $L$ be the total transform in $X$ of a line not passing through any of the points $p_1, \dots, p_r$. The classes of these divisors form a basis of $\text{Cl}(X)$; for convenience, we will write $e_i$ for the class $[E_i]$ of $E_i$ and $e_0$ for the class $[L]$. Further, the intersection product in $\text{Cl}(X)$ is defined by $e_i^2 = -1$ for $i=1, \dots, r$; $e_0^2 = 1$; and $e_i \cdot e_j = 0$ for all $i\neq j$. Let $Z_m = m(p_1+\cdots +p_r)$ be a uniform fat point subscheme with sheaf of ideals $\mathcal{I}_{Z_m}$; set $${F}_d = dE_0 - m(E_1 + E_2 + \cdots +E_r)$$ and $\mathcal{F}_d = \mathcal{O}_X(F_d)$. Then $\pi_{*}(\mathcal{F}_d) = \mathcal{I}_Z \otimes \mathcal{O}_{\mathbb{P}^2}(d)$ so $$\text{dim}((I_{Z_m})_d) = h^0(\mathbb{P}^2, \mathcal{I}_Z \otimes \mathcal{O}_{\mathbb{P}^2}(d)) = h^0(X, \mathcal{F}_d)$$ for all $d$. In particular, if $I \subseteq K[x,y,z]$ is the ideal of the points $p_1, \dots, p_r$ in $\mathbb{P}^2$, $$H_{I^{(m)}}(d) = h^0(X, \mathcal{F}_d)$$ and so we can study the Hilbert function of the symbolic powers $I^{(m)}$ by working with divisors on the surface $X$. For convenience, we will often write $h^0(X, F) = h^0(X, \mathcal{O}_X(F))$. Recall that if $[{F}]$ not the class of an effective divisor then $h^0(X, {F}) = 0$. On the other hand, if $F$ is effective, then we will see that we can compute $h^0(X,{F})$ by computing $h^0(X,{H})$ for some \textit{numerically effective} divisor $H$. \begin{defn} A divisor $H$ is \textbf{numerically effective} if $[F] \cdot [H] \geq 0$ for every effective divisor $F$, where $[F] \cdot [H]$ denotes the intersection multiplicity. The cone of classes of numerically effective divisors in $\text{Cl}(X)$ is denoted by NEF($X$). \end{defn} \begin{lem \label{lem:h0ofNEFF} Suppose that $X$ is the blow-up of $\mathbb{P}^2$ at $r \leq 8$ points in general position and that $F \in \text{NEF}(X)$. Then $F$ is effective and $$h^0(X, F) = ([F]^2-[F]\cdot [K_X])/2+1$$ where $K_X = -3e_0 + e_1 + \cdots + e_r$. \end{lem} \begin{proof} This is a consequence of Riemann-Roch and the fact that $h^1(X, F) = 0$ for any numerically effective divisor $F$. See Theorem 8 of \cite{Harbourne96} or Section 1 of \cite{FHH01} for a discussion. \end{proof} \begin{cor} \label{cor:h0ofFtNEFF} Let $F_t = tL - m(E_1+E_2 + \cdots + E_r)$. If $F_t$ is numerically effective then $$h^0(X,{F}_t) = {t+2 \choose 2}-r{m+1 \choose 2}.$$ \end{cor} A divisor class $[C]$ on $X$ is said to be \textit{exceptional} if it is the class of an exceptional divisor $C$ on $X$ (that is, a smooth curve isomorphic to $\mathbb{P}^1$ such that $[C]^2 = -1$).\footnote{Note that if $[C]$ is an exceptional class, there is a unique effective divisor in this class, typically called the \textit{exceptional curve}} The following result of Fichett, Harbourne, and Holay \cite{FHH01} tells us how to detect if a divisor is numerically effective if we know the exceptional curves. \begin{lem}[{Lemma 4(b) of \cite{FHH01}}] \label{lem:IdentifyingNEFF} Suppose that $X$ is the surface obtained by blowing up $2 \leq r \leq 8$ points of $\mathbb{P}^2$. Then $F$ is numerically effective if the intersection multiplicity of $[F]$ with all exceptional classes is greater than or equal to 0. \end{lem} Another result from \cite{FHH01} tells us what the exceptional curves of $X$ are in the cases that we are interested in. \begin{lem}[{Lemma 3(a) of \cite{FHH01}}] \label{lem:exceptionalcurves} Let $C$ be a curve on the blow-up $X$ of $\mathbb{P}^2$ at 8 points in general position. Then, with the notation above, the exceptional classes are the following, up to permutation of indices $1, 2, \dots, 8$: \begin{multicols}{2} \begin{itemize} \item $h_1=e_8$ \item $h_2=e_0-e_1-e_2$ \item $h_3=2e_0-e_1-\cdots -e_5$ \item $h_4=3e_0-2e_1-e_2-\cdots -e_7$ \item $h_5=4e_0-2e_1-2e_2-3e_3-e_4-\cdots -e_8$ \item $h_6=5e_0-2e_1-\cdots -2e_6-e_7-e_8$ \item $h_7=6e_0-3e_1-2e_2-\cdots -2e_8$. \end{itemize} \end{multicols} When $X$ is the blow-up of $\mathbb{P}^2$ at $n \leq 8$ points, the exceptional classes of $X$ are the ones listed above with $8-n$ of the $e_i$ ($i=1, \dots, 8$) set to 0. \end{lem} It turns out that knowing how to compute $h^0(X, H)$ for a numerically effective divisor $H$ will actually allow us to compute $h^0(X, F)$ for \textit{any} divisor $F$. In particular, for any divisor $F$, there exists a divisor $H$ such that $h^0(X, F) = h^0(X, H)$ and either: \begin{enumerate} \item[(a)] $H$ is numerically effective so $$h^0(X, F) = h^0(X, H) = (H^2-H\cdot K_X)/2+1$$ by Lemma \ref{lem:h0ofNEFF}; or \item[(b)] There is a numerically effective divisor $G$ such that $[G]\cdot [H] <0$ so $[H]$ is not the class of an effective divisor and $h^0(X, F) = h^0(X, H) = 0$. \end{enumerate} The following result will be used in Procedure \ref{proc:findH} to find such an $H$. \begin{lem} \label{lem:Equalh0} Suppose that $[C]$ is an exceptional class such that $[F] \cdot [C] <0$. Then $h^0(X, F) = h^0(X, F-C)$. \end{lem} \begin{proof} Note that it suffices to prove this statement for the case where $C$ is a smooth curve isomorphic to $\mathbb{P}^1$: if $[C']$ is an exceptional class then there exists a smooth curve $C$ isomorphic to $\mathbb{P}^1$ such that $[C'] = [C]$ so $[C'] \cdot [F] = [C] \cdot [F]$ and $h^0(X, F-C') = h^0(X, F-C)$. Note that we have an exact sequence $$ \mathcal{O}_X(F-C) \rightarrow \mathcal{O}_X(F) \rightarrow \mathcal{O}_C(F) \cong \mathcal{O}_{\mathbb{P}^1}([F] \cdot [C]) \rightarrow 0$$ induced by tensoring the exact sequence $$0 \rightarrow \mathcal{O}_X(-C) \rightarrow \mathcal{O}_X \rightarrow \mathcal{O}_C \rightarrow 0$$ with $\mathcal{O}_X(F)$. Then, from the long exact sequence of cohomology, $h^0(X, \mathcal{O}_X(F-C)) = h^0(X, \mathcal{O}_X(F))$ since $h^0(X, \mathcal{O}_{\mathbb{P}^1}([F]\cdot[C])) = 0$ ($[F]\cdot[C] <0$). \end{proof} The method for finding such the $H$ described above is as follows. \begin{procedure} \label{proc:findH} Given a divisor $F$ we can find a divisor $H$ with $h^0(X, F) = h^0(X, H)$ satisfying either condition (a) or (b) above as follows. \begin{enumerate} \item Reduce to the case where $[F] \cdot e_i \geq 0$ for all $i=1, \dots, n$: if $[F]\cdot e_i <0$ for some $i$, $h^0(X, F) = h^0(X, F-([F]\cdot e_i)E_i)$, so we can replace $F$ with $F - ([F]\cdot e_i) E_i$. \item Since $L$ is numerically effective, if $[F]\cdot e_0<0$ then $[F]$ is not the class of an effective divisor and we can take $H=F$ (case (b)). \item If $[F] \cdot [C] \geq 0$ for every exceptional class $[C]$ then, by Lemma \ref{lem:IdentifyingNEFF}, $F$ is numerically effective, so we can take $H = F$ (case (a)). \item If $[F] \cdot [C] <0$ for some exceptional class $[C]$ then $h^0(X, F) = h^0(X, F-C)$ by Lemma \ref{lem:Equalh0}. Then replace $F$ with $F-C$ and repeat from Step 2. \end{enumerate} \end{procedure} There are only a finite number of exceptional classes to check by Lemma \ref{lem:exceptionalcurves} so it is possible to complete Step 3. Further, it is easy to see with Lemma \ref{lem:exceptionalcurves} that $F \cdot e_0 > [F-C] \cdot e_0$ when $[C]$ is an exceptional curve, so the condition in Step 2 will be satisfied after at most $[F]\cdot e_0 +1$ repetitions. Thus, this process will eventually terminate.\footnote{The decomposition $F = H+(F-H)$ has been referred to as a \textit{Zariski decomposition} in some of the literature on fat points (for example, in \cite{FHH01}), but we avoid this terminology here because it is not consistent with definitions elsewhere (for example, in \cite{Lazarsfeld04}).} \subsection{Proof of Theorem \ref{thm:mainthm}(b)} \label{sec:678Proof} The proof of each part of Theorem \ref{thm:mainthm}(b) follows the same five steps outlined below. In Step 4, we will use the following lemma. \begin{lem} \label{lem:moninz} Let $I$ be the ideal of $r$ points in $\mathbb{P}^2$. The number of monomials in $\text{gin}(I^{(m)}) \subseteq K[x,y,z]$ of degree $t$ involving the variable $z$ is equal to $H_{I}(t-1)$. \end{lem} \begin{proof} Since, by Proposition \ref{prop:geninxy}, $\text{gin}(I^{(m)})$ is generated in the variables $x$ and $y$, the only elements of $\text{gin}(I^{(m)})_t$ that involve $z$ have to arise by multiplying monomials of $\text{gin}(I^{(m)})_{t-1}$ by $z$. Since multiplying each of the $H_{I^{(m)}}(t-1)$ monomials in $\text{gin}(I^{(m)})_{t-1}$ by $z$ gives distinct monomials, the result follows. \end{proof} As in Section \ref{sec:678Background}, $Z_m = m(p_1+\cdots +p_r)$ is a uniform fat point subscheme supported at $r$ distinct general points $p_1, \dots, p_r$ and $I$ is the ideal of $p_1, \dots, p_r$ so that $I^{(m)} = I_{Z_m}$. Recall that if $F_t = tL - m(E_1+\cdots +E_r)$ then $$H_{I_{Z_m}}(t) = h^0(X, \mathcal{F}_t);$$ we also write $H_{I_{Z_m}}(t) = H_Z(t)$. Finally, $$\alpha(m) := \min \{ t : H_{I_{Z_m}}(t) \neq 0\}.$$ \textbf{Step 1:} Find the smallest $N$ such that ${F}_t = tE_0-m(E_1+\cdots+E_r)$ is numerically effective for all $t \geq N$. To do this we will find the smallest $N$ such that $[F_t] \cdot [C] \geq 0$ for all $t \geq N$ (see Lemma \ref{lem:IdentifyingNEFF}). By Corollary \ref{cor:h0ofFtNEFF}, $$h^0(X, \mathcal{F}_t) = {t+2 \choose 2}-r{m+1 \choose 2}$$ for all $t \geq N$. \textbf{Step 2:} Use some optimal numerically effective divisor $D$ to find $M$ such that $[{F}_t]\cdot [D] <0$ for all $t<M$. By the definition of a numerically effective divisor, this will show that $[F_t]$ for $t<M$ is not the class of an effective divisor, and thus that $H_{I_{Z_m}}(t) = h^0(X, \mathcal{F}_t) = 0$ for all $t<M$. \textbf{Step 3:} Show that $h^0(X, \mathcal{F}_M) \neq 0$ where $M$ is as in Step 2. To do this, we will use Procedure \ref{proc:findH} to find a numerically effective $H$ such that $h^0(X, F_M) = h^0(X, H_M)$. Together with Step 2, this will show that $\alpha(m) = M$, and $x^M$ is the smallest power of $x$ in $\text{gin}(I^{(m)})$. \textbf{Step 4:} By Lemma \ref{lem:moninz}, the number of monomials of degree $t$ in the $\text{gin}(I_{Z_m})$ involving only the variables $x$ and $y$ is equal to $H_Z(t) -H_Z(t-1)$. Using this, show that the number of monomials in $\text{gin}(I_{Z_m})$ of degree $N+1$ in $x$ and $y$ is exactly $N+2$ (here $N$ is as in Step 1). This implies that all monomials in $x$ and $y$ of degree $N+1$ are in the $\text{gin}(I_{Z_m})$. Use Lemma \ref{lem:moninz} again to show that the number of monomials in $\text{gin}(I_{Z_m})$ of degree $N$ involving only $x$ and $y$ is strictly less than $N+1$, so not all monomials of degree $N$ in $x$ and $y$ are in $\text{gin}(I^{(m)})$. Since the ideals of the symbolic generic initial system are generated in $x$ and $y$ (Proposition \ref{prop:geninxy}), this will imply that $y^{N+1}$ is the smallest power of $y$ in $\text{gin}(I^{(m)})$. \textbf{Step 5:} The smallest power of $y$ in $\text{gin}(I^{(m)})$ is $N+1$ by Step 4 and the smallest power of $x$ in $\text{gin}(I^{(m)})$ is $\alpha(m)=M$ by Step 3. Thus, the intercepts of the limiting shape $P$ of the symbolic generic initial system of $I$ are $\big( 0, \lim_{m \rightarrow \infty} \frac{N+1}{m} \big)$ and $\big( \lim_{m \rightarrow \infty} \frac{M}{m}, 0 \big)$. Since $$ \Big( \lim_{m \rightarrow \infty} \frac{N+1}{m}\Big) \cdot \Big( \lim_{m \rightarrow \infty} \frac{M}{m} \Big) = r,$$ Corollary \ref{cor:InterceptsDetermineShape} implies that the limiting shape $P$ is as claimed in Theorem \ref{thm:mainthm}(b). \subsubsection{6 General Points} Throughout this section $I$ is the ideal of 6 points $p_1, \dots, p_6$ of $\mathbb{P}^2$ in general position and $Z_m = m(p_1 + \cdots + p_6)$ so $I_{Z_m} = I^{(m)}$. The exceptional classes of the blow-up $X$ of $\mathbb{P}^2$ at $p_1, \dots, p_6$ are those in Lemma \ref{lem:exceptionalcurves} with two of the $e_i$ set to 0. \textbf{Step 1:} To find an $N$ such that $F_t = tL-m(E_1+\cdots+E_6)$ is numerically effective for all $t \geq N$ we will use the permutation of the exceptional curves from Lemma \ref{lem:exceptionalcurves} that is most likely to make $h_i \cdot [F_t]$ negative. \begin{eqnarray*} F_t \cdot h_2= t - 2m \geq 0 &\iff& t \geq 2m\\ F_t\cdot h_3 = 2t - 5m \geq 0 &\iff& t \geq \frac{5}{2}m\\ F_t\cdot h_4 = 3t - 2m -5m\geq 0 &\iff& t \geq \frac{7}{3}m\\ F_t\cdot h_5 = 4t - 3\cdot 2m - 3m \geq 0 &\iff& t \geq \frac{9}{4}m\\ F_t \cdot h_6 = 5t - 2\cdot 6m \geq 0 &\iff& t \geq \frac{12}{5}m\\ F_t\cdot h_7 = 6t - 3m - 2\cdot 5 m \geq 0 &\iff& t \geq \frac{13}{6}m\\ \end{eqnarray*} The strongest condition on $t$ is $t \geq \frac{5}{2}m$. Thus, $N = \frac{5}{2}m$ and $F_t$ is numerically effective for all $t \geq \frac{5}{2}m$. Thus, $$H_{I^{(m)}}(t) = h^0(X, \mathcal{F}_t) = {t+2 \choose 2}-6{m+1 \choose 2}$$ for all $t \geq \frac{5}{2}m$. \textbf{Step 2:} Now we want to find an optimal numerically effective divisor $D$ such that $[{F}_t] \cdot [D]<0$ for small $t$. By the calculations in Step 1, $$D = 5L-2(E_1+\cdots +E_6)$$ is numerically effective ($D = F_5$ when $m=2$). If $[F_t]$ is the class of an effective divisor then $[D] \cdot[{F}_t] \geq 0$. Thus, if $[D] \cdot [{F}_t] <0$ then $[F_t]$ is not effective. Note that $$[D] \cdot [{F}_t] = 5t-2\cdot6m <0 \iff t< \frac{12m}{5}.$$ Thus, $[{F}_t]$ is not the class of an effective divisor when $t< \frac{12m}{5}$ so $h^0(X, \mathcal{F}_t) = 0$ for $ t< \frac{12}{5}m$. We set $M = \frac{12}{5}m$. \textbf{Step 3:} Starting with this step we will make a divisibility assumption on $m$. Suppose that $m$ is divisible by both 5 and 2, so $$m = 10m'$$ for some integer $m'$. The goal of this step is to show that $F_{\frac{12}{5}m} = F_{24m'}$ is in the class of an effective divisor; to do this we follow Procedure \ref{proc:findH}. One can check that the only exceptional class that has a negative intersection multiplicity with $[F_{24m'}]$ is $h_3=[2L-E_1- \cdots -E_5]$: $$[{F}_{24m'}] \cdot h_3 = 2 \cdot 24 m' - 5\cdot 10m' = -2m'.$$ At this point it will be useful to distinguish between the permutations of $h_3$; we will denote $[2L-E_1 - \cdots -\widehat{E_i} - \cdots - E_6]$ by $h_{3_i}$. \begin{lem} \label{lem:ifonethenall6} Let $F_t = tL-m(E_1+\cdots +E_6)$. If $[{F}_t] \cdot h_{3_1} <0$ then $([{F}_t] - h_{3_1} - \cdots - h_{3_i}) \cdot h_{3_{i+1}} <0$ for $i=1, \dots, 5$. \end{lem} \begin{proof} Suppose that $[{F}_t] \cdot h_{3_1}<0$, or equivalently, that $t< \frac{5}{2}m$. Then $$[{F}_t] - h_{3_1} - \cdots - h_{3_i}= (t-2i)e_0 - (m-(i-1))(e_1+\cdots +e_i) - (m-i)(e_{i+1} + \cdots + e_6)$$ so $$([{F}_t] - h_{3_1} - \cdots - h_{3_i})\cdot h_{3_{i+1}} = 2(t-2i) - (m-i+1)(i) - (m-i)(6-i-1) = 2t - 5m$$ which is less than 0 since $t< \frac{5}{2}m$. \end{proof} We will denote the sum $h_{3_1} + \cdots +h_{3_5}$ by $[Y]$ and call it a \textit{cycle}. Note that $$[Y]=(2\cdot 6)e_0 - 5(e_1+\cdots +e_6).$$ By Lemma \ref{lem:ifonethenall6}, if $[{F}_t] \cdot h_3 <0$ for one permutation then we subtract an entire cycle from $[F_t]$ when following Procedure \ref{proc:findH}. When following Procedure \ref{proc:findH}, we subtract off $2m'$ full cycles from $[{F}_{24m'}]$; \begin{eqnarray*} [{F}_{24m'}] - 2m'[Y] &=& (24m' -12 \cdot 2m')e_0 - (10m'-5\cdot 2m')(e_1+\cdots +e_6) \\ &=& 0e_0-0(e_1+\cdots +e_6) \end{eqnarray*} so $H_{24m'} = 0$. Therefore, $H_{I^{(m)}}(24m') = h^0(X, \mathcal{F}_{24m'}) = h^0(X, 0) = 1$ and $\alpha(m) = \frac{12m}{5}$ when $m$ is divisible by 10. \textbf{Step 4:} Again, in this section we will assume that 10 divides $m$ and we write $m = 10m'$ for some integer $m'$. Then $N=\frac{5}{2}m = 25m'$. By Lemma \ref{lem:moninz}, there are $H_Z(N+1) - H_Z(N)$ monomials in $\text{gin}(I^{(m)})$ that involve only $x$ and $y$. Since $F_N$ and $F_{N+1}$ are numerically effective by Step 1, \begin{eqnarray*} H_Z(N+1) - H_Z(N) &=& h^0(X,\mathcal{F}_{N+1}) - h^0(X,\mathcal{F}_{N})\\ &=& {N+2 \choose 1} = N+2. \end{eqnarray*} Thus, $\text{gin}(I_{Z_m})_{N+1}$ contains all monomials of degree $N+1$ in the variables $x$ and $y$. Now we need to determine $H_Z(N) - H_Z(N-1)$ and show that it is less than $N+1$ (that is, $\text{gin}(I_{Z_m})_N$ does not contain all monomials in $x$ and $y$ of degree $N$). Consider $$F_{N-1} = (25m'-1)L - 10m'(E_1+\cdots +E_6).$$ Then, following Procedure \ref{proc:findH}, we can subtract exactly 2 cycles $[Y]$ from $[{F}_{N-1}]$ to obtain $[H_{N-1}]$. We get $$[{H}_{N-1}] = (25m'-1-24)e_0 - (10m'-10)(e_1+ \cdots + e_6)$$ so, by Corollary \ref{cor:h0ofFtNEFF}, \begin{eqnarray*} h^0(X, \mathcal{F}_{N-1}) &=& h^0(X, \mathcal{H}_{N-1})\\ &=& \frac{25}{2}m'^2-\frac{35}{2}m'+6. \end{eqnarray*} By Step 1, $F_N$ is numerically effective so, again by Corollary \ref{cor:h0ofFtNEFF}, $$h^0(X, \mathcal{F}_N) = \frac{25}{2}m'^2+\frac{15}{2}m'+1$$ Thus, $$H_{I_{Z_m}}(N)-H_{I_{Z_m}}(N-1) = 25m'-5<N+1 = 25m'+1$$ and not all monomials in $x$ and $y$ of degree $N$ are contained in $\text{gin}(I^{(m)})$. Therefore, the largest degree generator of $\text{gin}(I_{Z_m})$ is of degree $N+1 = \frac{5}{2}m+1$ when $m$ is divisible by 10. \textbf{Step 5:} By Step 4, the highest degree generator of $\text{gin}(I^{(m)})$ is of degree $\frac{5}{2}m+1$ when $m$ is divisible by 10. By Step 3, the smallest degree element of $\text{gin}(I^{(m)})$ is of degree $\alpha(m) = \frac{12m}{5}$ when $m$ is divisible by 10. Thus, the intercepts of the limiting shape of the symbolic generic initial system of $I$ are $(0, \frac{5}{2})$ and $(\frac{12}{5},0)$. Since $$\frac{12}{5} \cdot \frac{5}{2} = 6,$$ Corollary \ref{cor:InterceptsDetermineShape} tells us that the boundary of the limiting shape is defined by the line through the intercepts and is as claimed in Theorem \ref{thm:mainthm}(b). \subsubsection{7 General Points} Throughout this section $I$ is the ideal of 7 points $p_1, \dots, p_7$ of $\mathbb{P}^2$ in general position and $Z = m(p_1 + \cdots + p_7)$ so $I_{Z_m} = I^{(m)}$. The exceptional classes of the blow-up $X$ of $\mathbb{P}^2$ at $p_1, \dots, p_7$ are those in Lemma \ref{lem:exceptionalcurves} with one of the $e_i$ set to 0. \textbf{Step 1:} To find an $N$ such that $F_t = tL-m(E_1+\cdots+E_7)$ is numerically effective for all $t \geq N$ we will use the permutation of the exceptional curves from Lemma \ref{lem:exceptionalcurves} that is most likely to make $h_i \cdot [F_t]$ negative. Similar to the case of six points, the strongest condition on $t$ from $h_i \cdot [F_t] \geq 0$ is $t \geq \frac{8}{3}m$. Thus, $N = \frac{8}{3}m$ and $F_t$ is numerically effective for all $t \geq N$. Further, $$H_{I^{(m)}}(t) = h^0(X, F_t) = {t+2 \choose 2}-7{m+1 \choose 2}$$ for all $t \geq \frac{8}{3}m$. \textbf{Step 2:} Now we want to find an optimal numerically effective divisor $D$. By the calculations in Step 1, $$D = 8L-3(E_1+\cdots +E_7)$$ is numerically effective ($D = F_8$ when $m=3$). If $[F_t]$ is the class of an effective divisor then $[D] \cdot [{F}_t] \geq 0$. We want to know when $[D] \cdot [{F}_t]$ is strictly less than 0 because this will imply that $[F_t]$ is not the class of an effective divisor. Note that $$[D] \cdot [{F}_t] = 8t-3\cdot7m <0 \iff t< \frac{21m}{8}.$$ Thus, $h^0(X, \mathcal{F}_t) = 0$ for $ t< \frac{21}{8}m$. We set $M = \frac{21}{8}m$. Our next goal is to show that this is an optimal value. That is, if $\frac{21}{8}m$ is an integer, then $h^0(X, \mathcal{F}_{\frac{21}{8}m}) \neq 0$. \textbf{Step 3:} Starting with this step we will make a divisibility assumption on $m$ and suppose that $m$ is divisible by both 8 and 3, so $$m = 24m'$$ for some integer $m'$. The goal of this step is to show that $F_{\frac{21}{8}m} = F_{21\cdot 3m'}$ is in the class of an effective divisor; to do this we will follow Procedure \ref{proc:findH}. One can check that the only exceptional class which has a negative intersection multiplicity with $[F_{63m'}]$ is $ h_4=[3L-2E_1- E_2-\cdots -E_7]$: $$[F_{21\cdot 3m'}] \cdot h_4 = 3 \cdot 63 m' -2\cdot 24m'-6\cdot 24m' = -3m'.$$ At this point it will be useful to distinguish between the permutations of $h_4$; we will denote $[3L-2E_i - E_1 - \cdots -\widehat{E_i} - \cdots - E_7]$ by $h_{4_i}$. \begin{lem} \label{lem:ifonethenall7} Let $F_t = tL-m(E_1+\cdots +E_7)$. If $[{F}_t] \cdot h_{4_1} <0$ then $([{F}_t] - h_{4_1} - \cdots - h_{4_i}) \cdot h_{4_{i+1}} <0$ for $i=1, \dots, 6$. \end{lem} We will denote the sum of all seven permutations of $h_4$ by $[Y]$ and call it one \textit{cycle}. Note that $$[Y]=21e_0 - 8(e_1+\cdots +e_7).$$ It is easy to see that we can subtract off $3m'$ full cycles from $[{F}_{21\cdot3m'}]$ to obtain $[H_{21\cdot3m'}]$. We have \begin{eqnarray*} [{F}_{21\cdot 3m'}] - 3m'[Y] &=& (21\cdot 3m' -21 \cdot 3m')e_0 - (24m'-8\cdot 3m')(e_1+\cdots +e_7)\\ &=& 0L-0(e_1+\cdots +e_7). \end{eqnarray*} so $H_{21\cdot3m'} = 0$ and $h_{I_{Z_m}}(21\cdot 3m') = h^0(X, F_{21\cdot 3m'}) = h^0(X, 0) = 1$ and $\alpha(m) = 21\cdot 3m' = \frac{21m}{8}$ when $m$ is divisible by $8\cdot3$. \textbf{Step 4:} Again, in this section we will assume that 24 divides $m$, so we write $m = 24m'$ for some integer $m'$. Then $N= \frac{8}{3}m = 8^2m'$. We now want to show that the number of monomials in the variables $x$ and $y$ in $\text{gin}(I^{(m)})_{N+1}$ is equal to $N+2$ and that the number of monomials of degree $N$ in $x$ and $y$ in $\text{gin}(I^{(m)})_N$ is less than $N+1$. This will prove that the highest degree generator occurs in degree $N+1$. By Lemma \ref{lem:moninz}, there are $H_Z(N+1) - H_Z(N)$ monomials in $\text{gin}(I^{(m)})$ that involve only $x$ and $y$. Then, since $F_N$ and $F_{N+1}$ are numerically effective by Step 1, $$H_Z(N+1) - H_Z(N) = {N+2 \choose 1} = N+2.$$ Thus, $\text{gin}(I_{Z_m})_{N+1}$ contains all monomials of degrees $N+1$ in the variables $x$ and $y$. Now we need to determine $H_Z(N) - H_Z(N-1)$ and show that it is less than $N+1$ (that is, $\text{gin}(I_{Z_m})_N$ does not contain all monomials in $x$ and $y$ of degree $N$). Consider $$F_{N-1} = (8^2m'-1)L - 24m'(E_1+\cdots +E_7).$$ Recall from Step 3 that one cycle is equal to $[Y]=3\cdot 7e_0 - 8(e_1+\cdots +e_7).$ By Procedure \ref{proc:findH}, $$[{H}_{N-1}] = (8^2m'-1-63)e_0 - (24m'-24)(e_1+ \cdots + e_7)$$ so, by Corollary \ref{cor:h0ofFtNEFF}, $$h^0(X, \mathcal{F}_{N-1}) = h^0(X, \mathcal{H}_{N-1}) = 32m'^2-52m'+21.$$ From Step 1 we know that ${F}_N$ is numerically effective and $$h^0(X, \mathcal{F}_N) = 32m'^2+12m'+1$$ Thus, $$H_{I_{Z_m}}(N)-H_{I_{Z_m}}(N-1) = 64m'-20<N+1 = 64m'+1$$ and not all monomials in $x$ and $y$ of degree $N$ are contained in $\text{gin}(I^{(m)})$. Therefore, the largest degree generator of $\text{gin}(I_{Z_m})$ is of degree $N+1 = \frac{8}{3}m+1$ when $m$ is divisible by 24. \textbf{Step 5:} By Step 4, the highest degree generator of $\text{gin}(I^{(m)})$ is of degree $\frac{8}{3}m+1$ when $m$ is divisible by 24. By Step 3, the smallest degree element of $\text{gin}(I^{(m)})$ is of degree $\alpha(m) = \frac{21m}{8}$ when $m$ is divisible by 24. Thus, the intercepts of the limiting shape of the symbolic generic initial system of $I$ are $(0, \frac{8}{3})$ and $(\frac{21}{8},0)$. Since $$\frac{8}{3} \cdot \frac{21}{8} = 7,$$ Corollary \ref{cor:InterceptsDetermineShape} tells us that the limiting polytope is defined by the line through the intercepts and is as claimed in Theorem \ref{thm:polytopefor678}. \subsubsection{8 General Points} Throughout this section $I$ is the ideal of 8 points $p_1, \dots, p_8$ of $\mathbb{P}^2$ in general position and $Z_m = m(p_1 + \cdots + p_8)$ so $I_{Z_m} = I^{(m)}$. The exceptional classes of the blow-up $X$ of $\mathbb{P}^2$ at $p_1, \dots, p_8$ are those in Lemma \ref{lem:exceptionalcurves}. \textbf{Step 1:} To find an $N$ such that $F_t = tL-m(E_1+\cdots+E_8)$ is numerically effective for all $t \geq N$ we compute $[C] \cdot [{F}_t]$ for all exceptional classes $[C]$ of $X$. Similar to the case of six points, we can check that the strongest condition on $t$ in resulting from $[C] \cdot [F_t] \geq 0$ is $t \geq \frac{17}{6}m$. Thus, $N = \frac{17}{6}m$ and $F_t$ is numerically effective for all $t \geq N$. Further, $$h^0(X, \mathcal{F}_t) = {t+2 \choose 2}-8{m+1 \choose 2}$$ for all $t \geq \frac{17}{6}m$. \textbf{Step 2:} Now we want to find an optimal numerically effective divisor $D$. By the calculations in Step 1, $$D = 17L-6(E_1+\cdots +E_8)$$ is numerically effective ($D = F_{17}$ when $m=6$). If $[F_t]$ is the class of an effective divisor then $[D] \cdot [{F}_t] \geq 0$. We want to know when $[D] \cdot [{F}_t]$ is strictly less than 0 which will imply that $[F_t]$ is not the class of an effective divisor. Note that $$[D] \cdot [{F}_t] = 17t-6\cdot8m <0 \iff t< \frac{48m}{17}.$$ Thus, $h^0(X, \mathcal{F}_t) = 0$ for $ t< \frac{48}{17}m$. We set $M = \frac{48}{17}m$. Our next goal is to show that this is an optimal value. That is, if $\frac{48}{17}m$ is an integer, then $h^0(X, \mathcal{F}_{\frac{48}{17}m}) \neq 0$. \textbf{Step 3:} Starting with this step we will make a divisibility assumption on $m$ and suppose that $m$ is divisible by both 17 and 6, so $$m = 17 \cdot 6m'$$ for some integer $m'$. The goal of this step is to show that $F_{\frac{48}{17}m} = F_{48\cdot 6m'}$ is in the class of an effective divisor. To do this we will follow Procedure \ref{proc:findH} to find $H_{48\cdot 5m'}$. One can check that the only exceptional class that has a negative intersection multiplicity with $[\mathcal{F}_{48 \cdot 6m'}]$ is $h_7=[6L-3E_1- 2E_2-\cdots -2E_8]$: $$[{F}_{48\cdot 6m'}] \cdot h_7 = 6 \cdot 48\cdot 6 m' -3\cdot 17\cdot 6m' - 2\cdot 7 \cdot 17\cdot 6 m' = -6m'.$$ It will be useful to distinguish between the permutations of $h_7$. We will denote $[6L-3E_i - 2E_1 - \cdots -\widehat{2E_i} - \cdots - 2E_7]$ by $h_{7_i}$. \begin{lem} \label{lem:ifonethenall8} Let $F_t = tL-m(E_1+\cdots +E_8)$. If $[{F}_t] \cdot h_{7_1} <0$ then $([{F}_t] - h_{7_1}- \cdots -h_{7_i}) \cdot h_{7_{i+1}} <0$ for $i=1, \dots, 7$. \end{lem} We will denote the sum of all eight permutations of $h_7$ by $[Y]$ and call it a \textit{cycle}. Note that $$Y=48e_0 - 17(e_1+\cdots +e_8).$$ Following Procedure \ref{proc:findH}, we subtract off $6m'$ full cycles from $[{F}_{48\cdot6m'}]$ to get $H_{48\cdot 6m'}$. Then \begin{eqnarray*} [F_{48\cdot 6m'}] - 6m'[Y] &=& (48\cdot 6m' -48 \cdot 6m')e_0 - (17\cdot 6m'-17\cdot 6m')(e_1+\cdots +e_8)\\ &=& 0e_0-0(e_1+\cdots +e_7). \end{eqnarray*} Therefore, $h^0(X, \mathcal{F}_{48\cdot 6m'}) = h^0(X, 0) = 1$ and $\alpha(m) = 48\cdot 6m' = \frac{48m}{17}$ in this case. \textbf{Step 4:} Again, in this section we will assume that 6 and 17 divide $m$, so $m = 17 \cdot 6m'$ for some integer $m'$ and $N = \frac{17}{6}m= 17^2m'$. We now want to show that the number of monomials in only $x$ and $y$ in $\text{gin}(I^{(m)})_{N+1}$ is $N+2$ and that the number of monomials of degree $N$ in the variables $x$ and $y$ in $\text{gin}(I^{(m)})$ is less than $N+1$. This will show that the highest degree generator occurs in degree $N+1$. By Lemma \ref{lem:moninz}, there are $H_Z(N+1) -H_Z(N)$ monomials in $\text{gin}(I^{(m)})$ that involve only $x$ and $y$. Then $$H_Z(N+1) - H_Z(N) = {N+2 \choose 1} = N+2$$ Thus, $\text{gin}(I_{Z_m})_{N+1}$ contains all monomials of degrees $N+1$ in the variables $x$ and $y$. Now we need to determine $H_Z(N) - H_Z(N-1)$ and show that it is less than $N+1$ (that is, $\text{gin}(I_{Z_m})_N$ does not contain all monomials in $x$ and $y$ of degree $N$). Consider $$F_{N-1} = (17^2m'-1)L - 17 \cdot 6m'(E_1+\cdots +E_8).$$ Recall that one cycle is equal to $[Y]=48e_0 - 17(e_1+\cdots +e_7).$ It is easy to see that exactly 6 cycles can be subtracted off of $[F_{N-1}]$ to obtain $[H_{N-1}]$. We have $$[H_{N-1}] = (17^2m'-1-6\cdot 48)e_0 - (17\cdot 6m'-17\cdot 6)(e_1+ \cdots + e_8)$$ so, by Corollary \ref{cor:h0ofFtNEFF}, $$h^0(X, \mathcal{F}_t) = \frac{289}{2}m'^2-\frac{527}{2}m'+120$$ From Step 1 we know that ${F}_N$ is numerically effective so $$h^0(X, \mathcal{F}_N) = \frac{289}{2}m'^2+\frac{51}{2}m'+1$$ Thus, $$H_Z(N)-H_Z(N-1) = 238m'-119<N+1 = 289m'+1$$ and not all monomials in $x$ and $y$ of degree $N$ are contained in $\text{gin}(I^{(m)})$. Therefore, the largest degree generator of $\text{gin}(I_{Z_m})$ is of degree $N+1 = \frac{17}{6}m+1$ when $m$ is divisible by $17\cdot 6$. \textbf{Step 5:} By Step 4, the highest degree generator of $\text{gin}(I^{(m)})$ is of degree $\frac{17}{6}m+1$ when $m$ is divisible by $17\cdot 6$. By Step 3, the smallest degree element of $\text{gin}(I^{(m)})$ is of degree $\alpha(m) = \frac{48m}{17}$ when $m$ is divisible by $17 \cdot 6$. Thus, the intercepts of the limiting shape of the symbolic generic initial system of $I$ are $(0, \frac{17}{6})$ and $(\frac{48}{17},0)$. Since $$\frac{17}{6} \cdot \frac{48}{17} = 8,$$ Corollary \ref{cor:InterceptsDetermineShape} tells us that the limiting polytope is defined by the line through the intercepts and is as claimed in Theorem \ref{thm:polytopefor678}. \section*{Acknowledgements} I thank Karen Smith for introducing this problem to me and for many useful discussions. I also thank Brian Harbourne and Susan Cooper for helping me to learn about fat points. \section{Preliminaries} \label{sec:prelim} In this section we will introduce some notation, definitions, and preliminary results related to fat points in $\mathbb{P}^2$, generic initial ideals, and systems of ideals. Unless stated otherwise, $R=K[x,y,z]$ is the polynomial ring in three variables over a field $K$ of characteristic 0 with the standard grading and some fixed term order $>$ with $x>y>z$. \subsection{Fat Points in $\mathbb{P}^2$} \label{sec:fatpoints} \begin{defn} Let $p_1, \dots, p_r$ be distinct points of $\mathbb{P}^2$, $I_j$ be the ideal of $K[\mathbb{P}^2] = R$ consisting of all forms vanishing at the point $p_j$, and $I = I_1 \cap \cdots \cap I_r$ be the ideal of the points $p_1, \dots, p_r$. A \textbf{fat point subscheme} $Z = m_1p_1 + \cdots + m_rp_r$, where the $m_i$ are nonnegative integers, is the subscheme of $\mathbb{P}^2$ defined by the ideal $I_Z = I_1^{m_1} \cap \cdots \cap I_r^{m_r}$ consisting of forms that vanish at the points $p_i$ to multiplicity at least $m_i$. When $m_i = m$ for all $i$, we say that $Z$ is \textbf{uniform}; in this case, $I_Z$ is equal to the $m^{\text{th}}$ symbolic power of $I$, $I^{(m)}$. \end{defn} The following lemma relates the symbolic and ordinary powers of $I$ in the case we are interested in (see, for example, Lemma 1.3 of \cite{AV03}). \begin{lem} \label{lem:symbpower} If $I$ is the ideal of distinct points in $\mathbb{P}^2$, $$(I^m)^{\text{sat}} = I^{(m)},$$ where $J^{\text{sat}} = \bigcup_{k \geq 0}(J:\mathfrak{m}^k)$ denotes the saturation of $J$. \end{lem} In this paper we will be interested in studying the ideals of uniform fat point subschemes $Z = mp_1 + \cdots + mp_r$ such that the points $p_1, \dots, p_r$ are in \textit{general position}. \begin{defn} A collection of points in $\mathbb{P}^2$ is in \textbf{general position} if, for each $d \in \mathbb{N}$, no subset of cardinality ${d+2 \choose 2}$ lies on any curve of degree $d$. \end{defn} \subsection{Generic Initial Ideals} An element $g = (g_{ij}) \in \text{GL}_n(K)$ acts on $R = K[x_1, \dots, x_n]$ and sends any homogeneous element $f(x_1, \dots, x_n)$ to the homogeneous element $$f(g(x_1), \dots, g(x_n))$$ where $g(x_i) = \sum_{j=1}^n g_{ij}x_j$. If $g(I)=I$ for every upper triangular matrix $g$ then we say that $I$ is \textit{Borel-fixed}. Borel-fixed ideals are \textit{strongly stable} when $K$ is of characteristic 0; that is, for every monomial $m$ in the ideal such that $x_i$ divides $m$, the monomials $\frac{x_jm}{x_i}$ are also in the ideal for all $j<i$. This property makes such ideals particularly nice to work with. To any homogeneous ideal $I$ of $R$ we can associate a Borel-fixed monomial ideal $\text{gin}_{>}(I)$ which can be thought of as a coordinate-independent version of the initial ideal. Its existence is guaranteed by Galligo's theorem (also see \cite[Theorem 1.27]{Green98}). \begin{thm}[{\cite{Galligo74} and \cite{BS87b}}] \label{thm:galligo} For any multiplicative monomial order $>$ on $R$ and any homogeneous ideal $I\subset R$, there exists a Zariski open subset $U \subset \text{GL}_n$ such that $\text{In}_{>}(g(I))$ is constant and Borel-fixed for all $g \in U$. \end{thm} \begin{defn} The \textbf{generic initial ideal of $I$}, denoted $\text{gin}_{>}(I)$, is defined to be $\text{In}_{>}(g(I))$ where $g \in U$ is as in Galligo's theorem. \end{defn} The \textit{reverse lexicographic order} $>$ is a total ordering on the monomials of $R$ defined by: \begin{enumerate} \item if $|I| =|J|$ then $x^I > x^J$ if there is a $k$ such that $i_m = j_m$ for all $m>k$ and $i_k < j_k$; and \item if $|I| > |J|$ then $x^I >x^J$. \end{enumerate} For example, $x_1^2 >x_1x_2 > x_2^2>x_1x_3>x_2x_3>x_3^2$. From this point on, $\text{gin}(I) = \text{gin}_{>}(I)$ will denote the generic initial ideal with respect to the reverse lexicographic order. Recall that the Hilbert function $H_I(t)$ of $I$ is defined by $H_I(t) = \text{dim}(I_t)$. The following theorem is a consequence of the fact that Hilbert functions are invariant under making changes of coordinates and taking initial ideals; we will use it frequently and freely throughout this paper. \begin{thm} \label{thm:commonproperties} For any homogeneous ideal $I$ in $R$, the Hilbert functions of $I$ and $\text{gin}(I)$ are equal. \end{thm} In this paper we will be studying the set of reverse lexicographic generic initial ideals of symbolic powers of a fixed ideal $I$, $\{\text{gin}(I^{(m)})\}_m$. One reason for our interest in these ideals is the following proposition which tells us that we can get information about the ideals $\text{gin}(I^{m})$ from the ideals $ \text{gin}(I^{(m)})$. \begin{prop}[{Proposition 2.21 of \cite{Green98}}] \label{prop:saturations} Fix the reverse lexicographic order on $K[x_1, \dots, x_n]$ with $x_1> x_2 > \cdots > x_n$ and let $\mathfrak{m} = (x_1, \dots, x_n)$. Then, if $I^{\text{sat}} = \bigcup_{k \geq 0}(I:\mathfrak{m}^k)$ denotes the saturation of $I$, $$\text{gin}(I^{\text{sat}}) = \bigcup_{k \geq 0}(\text{gin}(I):\mathfrak{m}^k) = (\text{gin}(I))^{\text{sat}}.$$ In particular, when $I$ is the ideal of distinct points in $\mathbb{P}^2$, $$\text{gin}(I^{(m)}) = \bigcup_{k \geq 0}(\text{gin}(I^m):\mathfrak{m}^k) = (\text{gin}(I^m))^{\text{sat}}.$$ for all $m \geq 1$ by Lemma \ref{lem:symbpower}. \end{prop} The following result due to Bayer and Stillman (\cite{BS87}). \begin{prop}[{Theorem 2.21 of \cite{Green98}}] \label{prop:geninxy} Fix the reverse lexicographic order on $K[x_1, \dots, x_n]$ with $x_1> x_2 > \cdots > x_n$. An ideal $I$ of $R$ is saturated if and only if no minimal generator of $\text{gin}(I)$ involves the variable $x_n$. In particular, when $I \subset K[x,y,z]$ is the (saturated) ideal of a set of distinct points of $\mathbb{P}^2$, no minimal generator of $\text{gin}(I^{(m)})$ involves the variable $z$. \end{prop} \begin{cor} \label{cor:formofgin} Suppose that $I \subset K[x,y,z]$ is the ideal of a set of distinct points of $\mathbb{P}^2$. Then the minimal generators of $\text{gin}(I^{(m)})$ under the reverse lexicographic order are of the form $$\{ x^{\alpha(m)}, x^{\alpha(m)-1}y^{\lambda_{\alpha(m)-1}(m)}, \dots, xy^{\lambda_1(m)}, y^{\lambda_0(m)} \}$$ where $\lambda_0(m) > \lambda_1(m) > \cdots > \lambda_{\alpha(m)-1}(m) \geq 1$. \end{cor} \begin{proof} By a result of Herzog and Srinivasan relating the dimension of a Borel-fixed monomial ideal $J$ to the variable powers that it contains, $\text{gin}(I^{(m)})$ contains a power of $y$ (see Lemma 3.1 of \cite{HS98}). Now the result is immediate from Proposition \ref{prop:geninxy} and the fact that $\text{gin}(I^{(m)})$ is a Borel-fixed ideal. \end{proof} \subsection{Graded Systems} In this subsection we introduce some tools for studying certain collections of monomial ideals. \begin{defn}[\cite{ELS01}] A \textbf{graded system of ideals} is a collection of ideals $J_{\bullet} =\{J_i\}_{i=1}^{\infty}$ such that $$J_i \cdot J_j \subseteq J_{i+j} \hspace{0.3in} \text{ for all } i, j \geq 1.$$ \end{defn} \begin{defn} The \textbf{generic initial system} of a homogeneous ideal $I$ is the collection of ideals $J_{\bullet}$ such that $J_i = \text{gin}(I^i)$. The \textbf{symbolic generic initial system} of a homogeneous ideal $I$ is the collection of ideals $J_{\bullet}$ such that $J_i = \text{gin}(I^{(i)})$. \end{defn} \begin{lem} The symbolic generic initial system is a graded system of ideals. \end{lem} \begin{proof} By definition, $\text{gin}(I^{(i)})$ is a monomial ideal. We need to show that for all $i,j \geq 1$, $\textnormal{gin}(I^{(i)})\cdot \textnormal{gin}(I^{(j)}) \subseteq \textnormal{gin}(I^{(i+j)})$. For any $l \geq 1$, let $U_l$ be the Zariski open subset of $GL_n$ such that $\text{gin}(I^{(l)})= \text{In}(g \cdot (I^{(l)}))$ for all $g$ in $U_l$. Since $U_i$, $U_j$, and $U_{i+j}$ are Zariski open they have a nonempty intersection; fix some $g \in U_i \cap U_j \cap U_{i+j}$. Given monomials $f' \in \textnormal{gin}(I^{(i)}) = \textnormal{In}(g(I^{(i)}))$ and $h' \in \textnormal{gin}(I^{(j)}) = \textnormal{In}(g(I^{(j)}))$, suppose that $f'=\textnormal{In}(g(f))$ and $h' = \textnormal{In}(g(h))$ for $f \in I^{(i)}$ and $h \in I^{(j)}$. Now $$f'\cdot h' = \textnormal{In}(g(f)) \textnormal{In}(g(h)) = \textnormal{In}(g(f) \cdot g(h)) = \textnormal{In}(g(f \cdot h)) \in \text{In}(g(I^{(i+j)}))$$ since $f \cdot h \in I^{(i+j)}$.\footnote{This holds since the set of symbolic powers of a fixed ideal is itself a graded system: $I^{(i)} \cdot I^{(j)} \subseteq I^{(i+j)}$.} Thus $f'\cdot h' \in \text{gin}(I^{(i+j)})$ as desired. \end{proof} The same proof with $I^{(i)}$ replaced by $I^i$ shows that the generic initial system is also a graded system of ideals. \begin{defn} [\cite{ELS03}] \label{defn:algebraicvolume} Let $\mathrm{a}_{\bullet}$ be a graded system of zero-dimensional ideals in $R=K[x_1, \dots, x_n]$. The \textbf{volume} of $\mathrm{a}_{\bullet}$ is $$\mathrm{vol}(\mathrm{a}_{\bullet}) := \limsup_{m \rightarrow \infty} \frac{n! \cdot \mathrm{length}(R/\mathrm{a}_m)}{m^n}.$$ \end{defn} Let $J$ be a monomial ideal of $R$. We may associate to $J$ a subset $\Lambda$ of $\mathbb{N}^n$ consisting of the points $\lambda$ such that $x^{\lambda} \in J$. The \textit{Newton polytope} $P_J$ of $J$ is the convex hull of $\Lambda$ regarded as a subset of $\mathbb{R}^n$. Scaling the polytope $P_J$ by a factor of $r$ gives another polytope which we will denote $rP_J$. If $\mathrm{a}_{\bullet}$ is a graded system of monomial ideals in $R$, the polytopes of $\{ \frac{1}{q} P_{\mathrm{a}_q} \}_q$ are nested: $\frac{1}{c}P_{\mathrm{a}_c} \subset \frac{1}{c+1}P_{\mathrm{a}_{c+1}}$ for all $c \geq 1$. The \textit{limiting shape $P$ of $\mathrm{a}_{\bullet}$} is the limit of the polytopes in this set: $$P = \bigcup_{q \in \mathbb{N}^*} \frac{1}{q} P_{\mathrm{a}_q}.$$ Under the additional assumption that the ideals of $\mathrm{a}_{\bullet}$ are zero-dimensional, the closure of each set $\mathbb{R}^n_{\geq 0} \backslash P_{\mathrm{a}_q}$ in $\mathbb{R}^n$ is compact. This closure is denoted by $Q_q$ and we let $$Q = \bigcap_{q \in \mathbb{N}^*} \frac{1}{q} Q_q.$$ \begin{prop} [\cite{Mustata02}] \label{prop:volumesequal} If $\textrm{a}_{\bullet}$ is a graded system of zero-dimensional monomial ideals in $R=K[x_1, \dots, x_n]$ and $Q$ is as defined above, $$\mathrm{vol}(\textrm{a}_{\bullet}) = n! \mathrm{ vol}(Q).$$ \end{prop} \begin{proof} This is an immediate consequence of Theorem 1.7 and Lemma 2.13 of \cite{Mustata02}. \end{proof} We now turn our attention to using the concept of the limiting shape to study the asymptotic behaviour of the system of ideals $\{ \text{gin}(I^{(m)})\}_m$ where $I$ is an ideal of $r$ distinct points in $\mathbb{P}^2$. By Corollary \ref{cor:formofgin}, the ideals $\text{gin}(I^{(m)})$ for such an $I$ are generated in the variables $x$ and $y$ and contain a power of both $x$ and $y$. Therefore, we can think of the ideals $\text{gin}(I^{(m)})$ as zero-dimensional in $K[x,y]$ and consider a two dimensional limiting shape $P$ of the symbolic generic initial system. \begin{lem} \label{lem:volume} Suppose that $I$ is the ideal of $r$ distinct points $p_1, p_2, \dots, p_r$ in $\mathbb{P}^2$ and $J_m = \text{gin}(I^{(m)}) \subseteq K[x,y]$. If $P$ is the limiting shape of $J_{\bullet}$ and $Q \subseteq \mathbb{R}^2$ is as above, $$\text{vol}(Q) = \frac{r}{2}.$$ \end{lem} \begin{proof} Let $h = ax+by+cz$ be a general linear form in $K[x,y,z]$. To reduce our calculations to $K[x,y]$, consider the ring isomorphism $$\phi: \frac{K[x,y,z]}{(h)} \rightarrow K[x,y]$$ given by sending $x$ to $x$, $y$ to $y$, and $z$ to $-\frac{a}{c}x - \frac{b}{c}y$. If $I_i\subseteq K[x,y,z]$ is the ideal of the point $p_i$ in $\mathbb{P}^2$ then $\phi(\overline{I_i}) \cong (x,y)^m$. Further, $\phi(\overline{I^{(m)}}) = \phi(\overline{I_1^m} \cap \cdots \cap \overline{I_r^m})$ and $\text{length} \Big( \frac{K[x,y]}{(x,y)^m} \Big) = {m+1\choose 2}$ so \begin{eqnarray*} \text{length} \bigg( \frac{K[x,y]}{\phi(\overline{I^{(m)}})}\bigg) &=& \text{length} \bigg( \frac{K[x,y]}{I_1^m} \times \cdots \times \frac{K[x,y]}{I_r^m} \bigg)\\ &=& \text{length} \bigg( \frac{K[x,y]}{(x,y)^m} \times \cdots \times \frac{K[x,y]}{(x,y)^m} \bigg)\\ &=&r(1+ \cdots +m). \end{eqnarray*} The fact that $\text{gin}(I^{(m)})$ is generated in $x$ and $y$ (Proposition \ref{prop:geninxy}) together with a well-known relation between the generic initial ideals of $J$ and $\phi(\overline{J})$ (see Corollary 2.5 of \cite{Green98}) imply that $\text{gin}(I^{(m)})$ and $\text{gin}(\phi(\overline{I^{(m)}}))$ have the same generators. Thus, thinking of $\text{gin}(I^{(m)})$ as being contained in $K[x,y]$, \begin{eqnarray*} \text{length} \bigg( \frac{K[x,y]}{\text{gin}(I^{(m)})}\bigg) &=& \text{length} \bigg( \frac{K[x,y]}{\text{gin}(\phi(\overline{I^{(m)}})}\bigg)\\ &=& \text{length} \bigg( \frac{K[x,y]}{\phi(\overline{I^{(m)}})}\bigg)\\ &=& r(1+\cdots +m) = r\Big( \frac{m^2+m}{2} \Big). \end{eqnarray*} Therefore, \begin{eqnarray*} \text{vol}(Q) &=& \lim_{m \rightarrow \infty} \frac{\text{length}(K[x,y]/\text{gin}(I^{(m)}))}{m^2}\\ &=& \lim_{m \rightarrow \infty} \frac{(m^2+m)r}{2m^2} \\ &=& \frac{r}{2}. \end{eqnarray*} \end{proof} If $I$ is the ideal of distinct points in $\mathbb{P}^2$, the minimal generating set of each ideal $\text{gin}(I^{(m)})$ contains a power of $x$ and a power of $y$, say $x^{\alpha(m)}$ and $y^{\zeta(m)}$ by Corollary \ref{cor:formofgin}. It is clear that $\lim_{m\rightarrow \infty} \frac{\alpha(m)}{m}$ and $\lim_{m\rightarrow \infty} \frac{\zeta(m)}{m}$ are the $x$- and $y$-intercepts of the limiting shape $P$ of $\{\text{gin}(I^{(m)})\}_m$. \begin{cor} \label{cor:InterceptsDetermineShape} Let $I \subseteq K[x,y,z]$ be the ideal of $r$ distinct points in $\mathbb{P}^2$ and $P$ be the limiting shape of the symbolic generic initial system $\{ \text{gin}(I^{(m)}) \}_m$. Suppose that the $x$-intercept $\gamma_1$ and the $y$-intercept $\gamma_2$ of the boundary of $P$ are such that $\gamma_1 \cdot \gamma_2 = r$. Then the limiting polytope $P$ has a boundary defined by the line passing through $(\gamma_1, 0)$ and $(0, \gamma_2)$. \end{cor} \begin{proof} The smallest possible limiting shape $P$ satisfying the given conditions is the one defined by the line segment through $(\gamma_1, 0)$ and $(0, \gamma_2)$ since $P$ is convex by definition. This extreme case is the only one in which the maximum volume under $P$ is achieved, in which case $\text{vol}(Q) = \frac{\gamma_1 \gamma_2}{2}$. Under the assumptions stated, $\gamma_1 \cdot \gamma_2 = r$ so, by the previous lemma, the maximum volume must be attained and $P$ is as claimed. \end{proof} \section{Introduction} Generic initial ideals can be viewed as a coordinate-independent version of initial ideals, which carry much of the same information as the initial ideal with the added benefit of preserving, and even revealing, certain geometric information. Given an ideal $I \subseteq K[x,y,z]$ of distinct points $p_1, \dots, p_r$ in $\mathbb{P}^2$, the reverse lexicographic generic initial ideal of $I$, $\text{gin}(I)$, can detect if a subset of the points lies on a curve of a certain degree (see \cite{EP90} or Theorem 4.4 of \cite{Green98}). If we instead consider the ideal $I^{(m)}$ of the fat point subscheme $Z_m = mp_1+\cdots +mp_r \subseteq \mathbb{P}^2$, one might ask what $\text{gin}(I^{(m)})$ says about $Z_m$; this question motivated the work in this paper. Despite being simple to describe, ideals $I^{(m)}$ of fat point subschemes $Z_m=m(p_1+\cdots+p_r)$ have proven difficult to understand. For example, there are still many open problems and unresolved conjectures related to finding the Hilbert function of $I^{(m)}$ and even the degree $\alpha(I^{(m)})$ of the smallest degree element of $I^{(m)}$. Many of the challenges in understanding the individual ideals $I^{(m)}$ can be overcome by changing one's focus to studying the general behaviour of the entire family of ideals $\{I^{(m)}\}_m$. For instance, more can be said about the Seshadri constant $$\epsilon(I) = \lim_{m \rightarrow \infty} \frac{\alpha(I^{(m)})}{rm}$$ than the invariants $\alpha(I^{(m)})$ of each ideal (see \cite{BC10} and \cite{Harbourne02} for further background on these constants). Thus, we will explore the asymptotic behaviour of the entire symbolic generic initial system $\{ \text{gin}(I^{(m)})\}_m$ as a first step to understanding the generic initial ideals of fat point subschemes. To describe limiting behaviour, we define the \textit{limiting shape} $P$ of the symbolic generic initial system $\{ \text{gin}(I^{(m)}\}$ of the ideal $I \subseteq K[x,y,z]$ corresponding to an arrangement of points in $\mathbb{P}^2$ to be the limit $\lim_{m \rightarrow \infty} \frac{1}{m} P_{\text{gin}(I^{(m)})}$, where $P_{\text{gin}(I^{(m)})}$ denotes the Newton polytope of $\text{gin}(I^{(m)})$. We will see that each of the ideals $\text{gin}(I^{(m)})$ is generated in the variables $x$ and $y$, so that $P_{\text{gin}(I^{(m)})}$, and thus $P$, can be thought of as a subset of $\mathbb{R}^2$. One reason for studying the limiting shape of a system of monomial ideals is that it completely determines the asymptotic multiplier ideals of the system (see \cite{Howald01} and \cite{Mayes12a}). When the point arrangement has an ideal $I$ that is a complete intersection of type $(\alpha, \beta)$ with $\alpha \leq \beta$, a special case of the main result of \cite{Mayes12a} shows that the limiting shape of the symbolic generic initial system has a boundary defined by the line through the points $(\alpha, 0)$ and $(0, \beta)$. The main result of this paper is the following theorem describing the limiting shape of the symbolic generic initial system of an ideal of $r$ distinct points of $\mathbb{P}^2$ in general position, assuming that the SHGH Conjecture \ref{conj:shgh} holds for the case where $r \geq 9$. \begin{thm} \label{thm:mainthm} Let $I \subseteq R=K[x,y,z]$ be the ideal of $r>1$ distinct points $p_1, \dots, p_r$ of $\mathbb{P}^2$ in general position and $P$ be the limiting shape of the reverse lexicographic symbolic generic initial system $\{ \text{gin}(I^{(m)})\}_m$. Then $P$ can be characterized as follows. \begin{enumerate} \item[(a)] If $r \geq 9$ and the SHGH Conjecture holds for infinitely many $m$, then $P$ has a boundary defined by the line through the points $(\sqrt{r},0)$ and $(0, \sqrt{r})$. See Figure \ref{fig:largenumberpolytope}. \item[(b)] If $6 \leq r<9$, then $P$ has a boundary defined by the line through the points $(\gamma_1,0)$ and $(0, \gamma_2)$ where: \begin{enumerate} \item[(i)] $\gamma_1 = \frac{12}{5}$ and $\gamma_2 = \frac{5}{2}$ when $r=6$; \item[(ii)] $\gamma_1 = \frac{21}{8}$ and $\gamma_2 = \frac{8}{3}$ when $r=7$; and \item[(iii)] $\gamma_1 = \frac{48}{17}$ and $\gamma_2 = \frac{17}{6}$ when $r=8$. \end{enumerate} \item[(c)] If $r=4$ or $r=5$, then $P$ has a boundary defined by the line through the points $(2,0)$ and $(0, \frac{r}{2})$. If $r=2$ or $r=3$, then $P$ has a boundary defined by the line through the points $(\frac{r}{2},0)$ and $(0,2)$. \end{enumerate} \end{thm} \begin{figure} \begin{center} \includegraphics[width=4cm]{GeneralPositionPolytopeBW.png} \end{center} \caption{The limiting shape $P$ of $\{\text{gin}(I^{(m)})\}_m$ where $I$ is the ideal of $r \geq 9$ points in general position, assuming that the SHGH Conjecture holds for infinitely many $m$.} \label{fig:largenumberpolytope} \end{figure} Precisely what information is carried by the limiting shape of the symbolic generic initial system of other point arrangements is still uncertain. While one can prove that the $x$-intercept of the boundary of $P$ is equal to $r \epsilon(I)$ (see Section \ref{sec:prelim}), that the $y$-intercept reflects the asymptotic behaviour of the regularity of the ideals $I^{(m)}$ (see \cite{Mayes12a}), and that the volume under $P$ is equal to $\frac{r}{2}$ (Proposition \ref{prop:volumesequal}), there is likely additional geometric information encoded within $P$. Two important questions concern the form of $P$: is $P$ always a polytope, and what does it mean for the boundary of $P$ to be defined by a certain number of line segments? Following background information in Section \ref{sec:prelim}, the three parts of Theorem \ref{thm:mainthm} are proven in Sections \ref{sec:largenumber}, \ref{sec:678}, and \ref{sec:smallnumber}. The final section contains an example demonstrating that there are point arrangements for which the boundary of the limiting polytope of the symbolic generic initial system is not defined by a single line segment. \section{The Symbolic Generic Initial System of Greater than 8 Uniform Points in General Position} \label{sec:largenumber} Throughout this section, $I \subseteq R[x,y,z]$ will denote the ideal of $r \geq 9$ points $p_1, \dots, p_r$ of $\mathbb{P}^2$ in general position. We will frequently use the fact that the Hilbert function of an ideal and its generic initial ideal are equal (see Theorem \ref{thm:commonproperties}). Computing the Hilbert functions of ideals of fat points in $\mathbb{P}^2$ can be very difficult. However, the following conjecture of Segre, Harbourne, Gimigliano, and Hirschowiz proposes that when $Z$ is the ideal of $r \geq 9$ uniform fat points in general position, $H_{I_Z}(t)$ has a very simple form. See \cite{Harbourne12} for a statement similar to what follows and \cite{Harbourne02} for more general versions of the conjecture. \begin{conj}[SHGH Conjecture] \label{conj:shgh} Let $R=K[x,y,z]$ and $I$ be the ideal of $r \geq 9$ generic points $p_i \in \mathbb{P}^2$. Then, if $I^{(m)}$ is the ideal of the uniform fat point subscheme $Z = m(p_1+\cdots+p_r)$, $$H_{I^{(m)}}(t) = \max \bigg \{ {t+2 \choose 2} - r {m+1 \choose 2}, 0 \bigg \}.$$ \end{conj} The SHGH Conjecture is known to hold for certain special cases. For example, it holds for infinitely many $m$ when $r$ is a square by \cite{HR04}, and for all $m$ when $r$ is a square not divisible by a prime bigger than 5 by \cite{Evain99}. The main goal of this section is to prove the first part of Theorem \ref{thm:mainthm}. \newtheorem*{thm:mainthma}{Theorem \ref{thm:mainthm}(a)} \begin{thm:mainthma} \label{thm:LargeAsymptoticBehaviour} Fix $r \geq 9$ points of $\mathbb{P}^2$ in general position and suppose that the SHGH Conjecture \ref{conj:shgh} holds for infinitely many $m$. Let $I$ be the ideal of $r$ general points in $\mathbb{P}^2$ and $P$ be the the limiting shape of the reverse lexicographic symbolic generic initial system $\{\text{gin}(I^{(m)})\}_m$. Then the boundary of $P$ is defined by the line through the points $(\sqrt{r},0)$ and $(0, \sqrt{r})$. \end{thm:mainthma} The proof of this statement is contained in Section \ref{sec:proofofLargeNumber}. In preparation for this proof, we compute the minimal generators of the generic initial ideals $\text{gin}(I^{(m)})$ in Section \ref{sec:StructureofLargeNumber} under the assumption that the SHGH Conjecture holds. \subsection{Structure of $\text{gin}(I^{(m)})$} \label{sec:StructureofLargeNumber} The following lemma records the degree of the smallest degree element of $I^{(m)}$. \begin{lem} \label{lem:alphavalue} Let $I$ be the ideal of $p_1, \dots, p_r$ points of $\mathbb{P}^2$ in general position where $r \geq 9$ and suppose that $\alpha(m)$ is the least integer $t$ such that $H_{I^{(m)}}(t) >0$. Then, if the SHGH Conjecture holds for $Z=m(p_1+\cdots p_r)$, $$\alpha(m) =\bigg \lfloor -\frac{1}{2} + \sqrt{\frac{1}{4} + rm^2+rm} \bigg \rfloor.$$ \end{lem} \begin{proof} By the SHGH Conjecture, $\alpha(m)$ is the smallest integer $t$ such that ${t+2 \choose 2} - r{m+1 \choose 2} > 0$. \begin{eqnarray*} &&\frac{(t+2)(t+1)}{2} - r\frac{(m+1)m}{2} >0\\ &\Leftrightarrow& t^2+3t+2-rm^2-rm > 0 \end{eqnarray*} If this is an equality, the positive root is $$t = - \frac{3}{2} + \frac{1}{2} \sqrt{1+4rm^2+4rm}.$$ Then the least integer that will make the expression positive is $$\alpha(m) = \bigg \lfloor -\frac{3}{2} + \frac{1}{2}\sqrt{1 + 4rm^2+4rm} +1 \bigg \rfloor.$$ \end{proof} If the SHGH Conjecture holds, the structure of the generic initial ideals $\text{gin}(I^{(m)})$ is very simple. \begin{prop} \label{prop:gensofgin} Let $I$ be the ideal of $r \geq 9$ points of $\mathbb{P}^2$ in general position, fix a non-negative integer $m$, and suppose that the SHGH Conjecture holds for $I^{(m)}$. Set $\alpha = \alpha(m)$ and $\eta := H_{I^{(m)}}(\alpha) = {\alpha+2 \choose 2} - r{m+1 \choose 2}$ so that $\eta \leq \alpha+1$. Then $$\text{gin}(I^{(m)}) = (x^{\alpha}, x^{\alpha-1}y, \dots, x^{\alpha-\eta+1}y^{\eta-1}, x^{\alpha-\eta}y^{\eta+1}, x^{\alpha-\eta-1}y^{\eta+2}, \dots, xy^{\alpha}, y^{\alpha+1})$$ when $\eta < \alpha+1$ and $$\text{gin}(I^{(m)}) = (x^{\alpha}, x^{\alpha-1}y, \dots, xy^{\alpha-1}, y^{\alpha})$$ when $\eta = \alpha+1$. \end{prop} \begin{proof} Since there is no element of $\text{gin}(I^{(m)})$ of degree smaller than $\alpha(m)$, all monomials of degree $\alpha(m)$ in $\text{gin}(I^{(m)})$ must be generators and thus contain only the variables $x$ and $y$ by Proposition \ref{prop:geninxy}. There are at most $\alpha+1$ monomials of degree $\alpha$ in the variables $x$ and $y$ so $\eta := H_{\text{gin}(I^{(m)})}(\alpha) \leq \alpha+1$. If $\eta = \alpha+1$ then all $\alpha+1$ monomials of degree $\alpha$ in the variables $x$ and $y$ are minimal generators of $\text{gin}(I^{(m)})$. By Corollary \ref{cor:formofgin}, $\text{gin}(I^{(m)})$ has exactly $\alpha+1$ minimal generators. Thus, all minimal generators of $\text{gin}(I^{(m)})$ are of degree $\alpha$ and are the ones given. Now suppose that $\eta < \alpha+1$. The $\eta$ monomials of $\text{gin}(I^{(m)})$ of degree $\alpha$ must be minimal generators. In fact, since generic initial ideals are Borel-fixed, these must be the largest $\eta$ monomials in $x$ and $y$ of degree $\alpha$ with respect to the reverse lexicographic order: $$ [\text{gin}(I^{(m)})]_{\alpha}= \{ x^{\alpha}, x^{\alpha-1}y, \dots, x^{\alpha-\eta+1}y^{\eta-1} \}.$$ There are exactly $\eta$ elements of $\text{gin}(I^{(m)})$ of degree $\alpha+1$ involving the variable $z$, obtained by multiplying each of the $\eta$ generators of $[\text{gin}(I^{(m)})]_{\alpha}$ by $z$. By the SHGH Conjecture \ref{conj:shgh}, \begin{eqnarray*} H_{I^{(m)}}(\alpha+1) - \eta &=& \bigg[ {\alpha+1+2 \choose 2} - r{m+1\choose 2} \bigg] - \bigg[ {\alpha+2 \choose 2} - r{m+1\choose 2} \bigg]\\ &=& {\alpha+2+1 \choose 2} - {\alpha+2 \choose 2} \\ &=& {\alpha+2 \choose 1} = \alpha+2 \end{eqnarray*} and there are $\alpha+2$ monomials in $\text{gin}(I^{(m)})$ of degree $\alpha+1$ containing only the variables $x$ and $y$. Since there are exactly $\alpha+2$ monomials of degree $\alpha+1$ in $x$ and $y$, $\text{gin}(I^{(m)})$ contains all of them. Thus, the remaining generators of $\text{gin}(I^{(m)})$ are of degree $\alpha+1$; they are $$x^{\alpha-\eta}y^{\eta+1}, x^{\alpha-\eta-1}y^{\eta+2}, \dots, xy^{\alpha}, y^{\alpha+1}$$ by Corollary \ref{cor:formofgin}. \end{proof} \subsection{Proof of Theorem \ref{thm:mainthm} (a)} \label{sec:proofofLargeNumber} \begin{proof}[Proof of Theorem~\ref{thm:mainthm} (a)] By Proposition \ref{prop:gensofgin}, $x^{\alpha(m)}$ and $y^{\alpha(m)+1}$ or $y^{\alpha(m)}$ are the smallest variable powers contained in $\text{gin}(I^{(m)})$ for all $m$ such that the SHGH Conjecture holds. Thus, the $x$-intercept of the boundary of $P$ is $$\lim_{m \rightarrow \infty} \frac{\alpha(m)}{m}$$ while the $y$-intercept of the boundary of $P$ is $$\lim_{m \rightarrow \infty} \frac{\alpha(m)+1}{m} = \lim_{m \rightarrow \infty} \frac{\alpha(m)}{m}$$ where we take the limits over the infinite subset such that the SHGH Conjecture holds. By Lemma \ref{lem:alphavalue}, \begin{eqnarray*} \lim_{m \rightarrow \infty} \frac{\alpha(m)}{m} &=& \lim_{m \rightarrow \infty} \frac{\big \lfloor -\frac{1}{2} + \sqrt{\frac{1}{4} + rm^2+rm} \big \rfloor}{m}\\ &=& \sqrt{r} \end{eqnarray*} so the $x$ and $y$ intercepts of the limiting shape $P$ are both equal to $\sqrt{r}$. Since $\sqrt{r} \cdot \sqrt{r} = r$, Corollary \ref{cor:InterceptsDetermineShape} tells us that the boundary of $P$ is defined by the line through the $x$- and the $y$-intercepts as claimed. \end{proof} \section{The Symbolic Generic Initial System of 5 or Fewer Uniform Fat Points in General Position} \label{sec:smallnumber} In this section we prove part (c) of the main theorem. \newtheorem*{thm:mainthmc}{Theorem \ref{thm:mainthm} (c)} \begin{thm:mainthmc} \label{thm:ShapeofFiveorLess} Suppose that $1<r \leq 5$ and $I$ is the ideal of $r$ points in general position. Then the limiting polytope of the symbolic generic initial system $\{\text{gin}(I^{(m)})\}$ has a boundary defined by the line through the points $(2,0)$ and $(0, \frac{r}{2})$ when $\frac{r}{2} \geq 2$ and $(\frac{r}{2},0)$ and $(0, 2)$ when $\frac{r}{2}<2$. \end{thm:mainthmc} This is an immediate consequence of the following result of \cite{Mayes12d} since five or fewer points of $\mathbb{P}^2$ in general position lie on an irreducible conic. \begin{thm} Suppose that $I$ is the ideal of $r$ points in $\mathbb{P}^2$ lying on an irreducible conic. Then the limiting polytope of the symbolic generic initial system $\text{gin}(I^{(m)})$ has a boundary defined by the line through the points $(2,0)$ and $(0, \frac{r}{2})$ when $\frac{r}{2} \geq 2$ and $(\frac{r}{2},0)$ and $(0, 2)$ when $\frac{r}{2}<2$. \end{thm} \section{Final Example} The results presented here and in \cite{Mayes12a} may lead the reader to believe that the limiting polytope of any symbolic generic initial system is defined by a single hyperplane. The following example shows that this does not hold even for ideals of points in $\mathbb{P}^2$. \begin{example} \label{example:DifferentShape} Suppose that $I$ is the ideal of the $l+1 \geq 4$ points $p_1, \dots, p_l, p_{l+1}$ of $\mathbb{P}^2$ where $p_1, \dots p_l$ lie on a line and $p_{l+1}$ lies off of the line. \begin{prop} \label{prop:lpointson} Let $I$ be the ideal of $l+1$ distinct points of $\mathbb{P}^2$ where $l$ of the points lie on a line and suppose that $l(l-1)$ divides $m$. Then the highest degree generator of $\text{gin}(I^{(m)})$ is of degree $lm$ and the lowest degree generator of $\text{gin}(I^{(m)})$ is of degree $2m-\frac{m}{l}$. \end{prop} \begin{proof}[{Idea of Proof}] The proof of this proposition is similar to the work contained in Section \ref{sec:678} with the following considerations. In this case, the blow-up $\pi: X \rightarrow \mathbb{P}^2$ of $p_1, \dots, p_{l+1}$ has exceptional curves with classes $[L-E_1-E_2 - \cdots -E_l]$ and $[L-E_i-E_{l+1}]$ for $i=1, \dots, l$ where $E_j = \pi^{-1}(p_j)$ and $L$ is the total transform of a general line in $\mathbb{P}^2$ (note that the exceptional curves are the total transforms of lines through the points $\mathbb{P}^2$; see \cite{Harbourne98}). \end{proof} If $P$ is the limiting shape of the symbolic generic initial system $\{ \text{gin}(I^{(m)}) \}_m$, then Proposition \ref{prop:lpointson} implies that the boundary of $P$ has $y$-intercept $$\lim_{ m \rightarrow \infty} \frac{lm}{m} = l$$ and $x$-intercept $$\lim_{m \rightarrow \infty} \frac{2m- \frac{m}{l}}{m} = 2- \frac{1}{l}.$$ If the boundary of $P$ was defined by the line through these intercepts, the volume under of $P$ would be $$\text{vol}(Q) = \frac{(l)(2-\frac{1}{l})}{2} = l -\frac{1}{2}.$$ However, by Lemma \ref{lem:volume}, the volume under of $P$ must be $\frac{l+1}{2}$ which is strictly smaller than $l-\frac{1}{2}$ ($l \geq 3$). Thus, $P$ is not defined by the line through the intercepts. In fact, one can prove that the limiting polytope $P$ is the one shown in Figure \ref{fig:lpointsonpolytope}. \end{example} \vspace{0.2in} \begin{figure} \begin{center} \includegraphics[width=4.5cm]{OtherShapeBW.png} \end{center} \caption{The limiting polytope $P$ of the symbolic generic initial system of the ideal of $l$ points on a line and one point off.} \label{fig:lpointsonpolytope} \end{figure}
\section{Introduction} \label{sec:introduction} With the discovery of what is expected to be the Higgs particle \cite{higgsmass}, the Standard Model of particle physics now provides a coherent and consistent theory of fundamental physics up to and including the electroweak scale. Although many phenomena in the realm of cosmology, such as inflation, dark matter and dark energy, are not addressed in this framework, all experimentally observed processes are very well described\footnote{We will consider the Standard Model to include right-handed neutrinos and a non-zero mass term for neutrinos. Hence the phenomenon of neutrino oscillations is considered part of the Standard Model.}. One central issue at the boundary between cosmology and particle physics is the origin of the baryon asymmetry observed in the Universe. A substantial effort has been made to link this phenomenon to electroweak scale physics \cite{EWBG1,EWBG2,EWBG3}, since it is the lowest energy where baryon number violation may occur. In combination with C-, P- and CP-violation and an out-of-equilibrium electroweak symmetry breaking transition, baryon asymmetry with the observed magnitude can indeed be produced. However, because Standard Model CP-violation is minute \cite{shaposhCP} (see however \cite{CPV1}) and the electroweak transition is a crossover in the Standard Model for the physical Higgs mass of 125-126 GeV \cite{rummukainen}, electroweak baryogenesis requires additional fields and interactions to exist. Presumably the simplest way to achieve successful electroweak baryogenesis is to extend the scalar sector of the Standard Model by an additional field. This could be a $SU(2)$ singlet but probably the most popular extension is the Two-Higgs Doublet Model (2HDM) with an additional $SU(2)$ doublet. A number of different ``types'' exist, depending on how the two Higgs doublets couple to the fermions (see \cite{2hdmreview} for a recent review). The most general Higgs potential contains 14 (real) parameters, including up to two CP violating phases. In addition, because of the richness of the vacuum structure CP may be spontaneously broken. Restricting to a sub-class of models with only 10 real parameters, we ask the question whether the observed baryon asymmetry can be used to constrain the parameter space, complementing direct collider experiments. We expect that masses (4 different, of which we know the lightest) and vevs (2, of which we know one) are the easiest to measure, and so we will imagine that in future these are constrained, leaving a 4-dimensional less accessible subspace. Our aim here is to show how one may in principle sweep through this subspace and potentially use the observed baryon asymmetry to pin down the allowed parameter region. The 2HDM doublet can accommodate a strong first order phase transition, but we will consider a different scenario, where electroweak symmetry breaking is a cold spinodal transition \cite{CEWBAG1,CEWBAG2,CEWBAG3,CEWBAG4}. This is a viable alternative to the standard ``Hot'' scenario \cite{EWBG2}, but for our purpose here, its main virtue is that it is straightforward to simulate numerically from first principles. Cold electroweak baryogenesis may be realized as a result of coupling to another scalar field, which may \cite{inflation} or may not \cite{notinflation,servant} be the inflaton. We demonstrated in \cite{paper01} through direct numerical simulations that a baryon asymmetry is indeed produced, as a result of the interplay between the explicit CP/C violation in the Higgs potential and the C- and P-violating (but CP-conserving) gauge-fermion interactions. It turns out that when simulating a bosonized version of the theory, it is necessary to include the P-breaking of fermions, and we did this through an effective higher order bosonic interaction term, parameterized by a coefficient $\delta_{\rm C/P}$. Although this coefficient can in principle be computed analytically, this is a very non-trivial task and we chose to keep it as a free parameter. We found that in order for the observed asymmetry to be reproduced, we need $\delta_{\rm C/P}\simeq (\textrm{2 to 3})\times 10^{-4}$ or larger. The paper is structured as follows: In section \ref{sec:model} is a brief introduction of the bosonized 2HDM. In section \ref{sec:alpha} we parametrize the 4-dimensional parameter space in terms of a field transformation, two angles and one mass scale. The numerical results are presented in section \ref{sec:fullnum}, where we for a given range of parameters, and using lattice simulations in real-time, directly compute the baryon asymmetry in the bosonized electroweak sector. We conclude in section \ref{sec:conclusion}. In Appendix \ref{app:higgs}, we further discuss the parametrization of the neutral Higgs masses. \section{The 2HDM} \label{sec:model} The 2HDM is defined through the continuum action \begin{eqnarray} \label{eq:action} S=-\int d^3x\, dt\,\bigg[ \frac{1}{4g^2}\textrm{Tr}\, F_{\mu\nu}F^{\mu\nu}+ (D_\mu\phi_1)^\dagger D^\mu\phi_1+(D_\mu\phi_2)^\dagger D^\mu\phi_2+V(\phi_1,\phi_2) +\mathcal{L}_\textrm{C/P} \bigg],\nonumber\\ \end{eqnarray} where we use the metric $\eta={\rm diag}(-+++)$, $\phi_{1,2}$ are SU(2) doublets with hypercharge $+1$ and $F^{\mu\nu}$ is the field strength tensor of the gauge field. We will ignore the SU(3) and U(1) gauge fields. The covariant derivative is $D_\mu\phi_i=(\partial_\mu+iA_\mu)\phi_i$ and the potential is in all generality \footnote{We here correct an error in \cite{paper01} in the normalization of the coefficients. The results obtained there were based on the conventions presented here.} \begin{eqnarray}\label{equ:V} V(\phi_1,\phi_2)&=&-\frac{\mu_{11}^2}{2}\phi_1^\dagger\phi_1-\frac{\mu_{22}^2}{2}\phi_2^\dagger\phi_2-\frac{\mu^2_{12}}{2}\,\phi_1^\dagger\phi_2-\frac{\mu^{2,*}_{12}}{2}\phi_2^\dagger\phi_1\nonumber\\ &&+\frac{\lambda_1}{2}(\phi_1^\dagger\phi_1)^2+\frac{\lambda_2}{2}(\phi_2^\dagger\phi_2)^2+\lambda_3(\phi_1^\dagger\phi_1)(\phi_2^\dagger\phi_2)+\lambda_4(\phi_2^\dagger\phi_1)(\phi_1^\dagger\phi_2)\nonumber\\ &&+\frac{\lambda_5}{2}(\phi_1^\dagger\phi_2)^2+\frac{\lambda_5^*}{2}(\phi_2^\dagger\phi_1)^2 +\lambda_6(\phi_1^\dagger\phi_1)(\phi_1^{\dagger}\phi_2) +\lambda_6^*(\phi_1^\dagger\phi_1)(\phi_2^{\dagger}\phi_1)\nonumber\\ &&+\lambda_7(\phi_2^\dagger\phi_2)(\phi_1^{\dagger}\phi_2) +\lambda_7^*(\phi_2^\dagger\phi_2)(\phi_2^{\dagger}\phi_1). \end{eqnarray} The parameters $\lambda_{1,2,3,4}$ and $\mu^2_{11,22}$ are real and in general $\lambda_{5,6,7}$ and $\mu_{12}^2$ are complex. In this paper, we only study the 2HDM with a softly broken $Z_2$ symmetry, in which $\lambda_6=\lambda_7=0$ \cite{2hdmreview}. There is then only one independent CP violating phase. In the Standard Model as well as the 2HDM there is also CP-violation through the complex phase in the CKM mixing matrix. For the purpose of the present work, we will assume that the effective CP-breaking terms arising from this are negligible, although at very low temperatures, this may not be correct \cite{CPV1}. We will take (\ref{eq:action}) to represent a bosonized version of the full theory, where fermions have been integrated out, and their effect is captured in a C- and P-breaking term given by \cite{turok} \begin{eqnarray} \label{eq:CandP} \mathcal{L}_{\rm C/P}=\frac{\delta_{\rm C/P}}{16\pi^2 m_W^2}i(\phi_1^\dagger \phi_2-\phi_2^\dagger \phi_1)\textrm{ Tr}\,F_{\mu\nu}\tilde{F}^{\mu\nu}, \end{eqnarray} The Yukawa couplings and the mixing matrix is encoded in the real parameter $\delta_{\rm C/P}$, and it can in principle be computed from the model. The standard prescription in bosonized theories, which we will also adopt here, is then to infer the value of the baryon number $B$ through the anomaly equation \begin{eqnarray} \label{eq:anomalyeq} B(t)-B(0)=3[N_{\rm cs}(t)-N_{\rm cs}(0)], \end{eqnarray} where $N_{\rm cs}$ is the Chern-Simons number of the SU(2) gauge field. The reason for including the term (\ref{eq:CandP}) is that, as demonstrated in \cite{paper01}, to generated a non-zero average Chern-Simons number, we need P-symmetry to be broken as well as CP-symmetry. It is easy to see that (\ref{eq:CandP}) conserves CP. It turns out that in a finite temperature environment, the Higgs winding numbers $N_{\rm w}^{1,2}$ for the two Higgs fields, respectively, are much cleaner observables. At late times, the three agree, $N_{\rm w}^{1,2}=N_{\rm cs }$, and so we will identify the winding numbers at the end of the simulation to be the late time value for Chern-Simons number and hence the baryon asymmetry. \section{Choices of the parameters} \label{sec:alpha} \subsection{The full parameter space of the 2HDM} \label{sec:fullparam} We will re-parametrize the 10-dimensional parameter space in the following way: \begin{itemize} \item {\bf Vacuum parameters (3):} $v$, $\beta$ and $\theta$.\\ Without loss of generality, we can parametrize the Higgs fields in terms of 2 complex and 4 real degrees or freedom as \begin{equation} \phi_1= e^{i\theta}\left(\begin{array}{c} \phi_1^+\\ \left( v_1 + \eta_1 + i \chi_1\right)/\sqrt{2} \end{array}\right),\qquad\phi_2=\left(\begin{array}{c} \phi_2^+\\ \left( v_2 + \eta_2 + i \chi_2\right)/\sqrt{2} \end{array}\right). \end{equation} and \begin{eqnarray} \chi_1 &=& \cos\beta G^0 - \sin\beta \eta^3,~~\chi_2 = \cos\beta \eta^3 + \sin\beta G^0,\\ \phi_1^+ &=& \cos\beta G^+ - \sin\beta H^+,~~\phi_2^+ = \cos\beta H^+ + \sin\beta G^+. \end{eqnarray} The vacuum is given by $G^{0,+}=\phi_{1,2}^+=\eta_{1,2,3}=0$, in terms of $v_1e^{i\theta}$ and $v_2$. We introduce $v$ and $\beta$ through \begin{eqnarray} v_1 = v \cos\beta, \qquad v_2=v \sin\beta, \qquad v_2/v_1=\tan\beta. \end{eqnarray} Minimizing the Higgs potential gives three equations \begin{equation} \frac{\partial}{\partial v_1} V|_{v_1,v_2,\theta}=0,\quad \frac{\partial}{\partial v_2} V|_{v_1,v_2,\theta}=0,\quad \frac{\partial}{\partial \theta} V|_{v_1,v_2,\theta}=0, \end{equation} with which we can replace three couplings/mass parameters by $\beta$, $\theta$ and $v$. \item {\bf Higgs masses (4) :} $m_{1,2,3}$ and $m_\pm$.\\ There are five physical Higgs bosons: two form one charged field $H^\pm$ and the rest are mass eigenstates formed as linear combinations of the neutral fields $\eta_{1,2,3}$. We introduce the mass eigenvalues for these, $m_\pm$ and $m_{1,2,3}$, respectively, and these replace four other parameters (see also Appendix \ref{app:higgs}). \item {\bf Neutral Higgs mixing angles (2) :} $\alpha_1$, $\alpha_2$.\\ As discussed in Appendix \ref{app:higgs}, the mass matrix of the neutral Higgs modes is in general not diagonal in the fields $\eta_{1,2,3}$, but it can be diagonalized through three mixing angles $\alpha_{1,2,3}$. Only two of these are independent, and we take $\alpha_3$ to be fixed through Eq.~(\ref{eq:alpha3}), which has 0, 1 or 2 solutions for a given set of $(\alpha_{1},\alpha_2)$. \item {\bf A mass parameter (1) :} $\mu^2=\textrm{Re}(\mu_{12}^2 e^{-i\theta})$. \end{itemize} At the end of the day, the parameter set in the Higgs potential denoted by $\{\lambda\}$ is a function of the above 10 parameters. In the following discussion, for simplicity of notation we will use $\{\lambda\}[\ldots]$ with the ellipsis being some of the above parameters relevant for the discussions only. \subsection{The subspace spanned by ($\alpha_{1}$, $\alpha_{2}$, $\theta$, $\mu$)} \label{sec:4param} As explained in the Introduction, we will assume that the 4 distinct masses and the two vevs have been determined (or at least constrained) by experiment, so that we can assign values to them: \begin{itemize} \item {\bf Vacuum parameters:}\\ The vev $v$ is known but not $\beta$ and we choose \begin{eqnarray} v = 246~\text{GeV}, \qquad \tan\beta = 2.\label{equ:vac} \end{eqnarray} \item {\bf Higgs masses:} The lowest neutral Higgs mass $m_1$ is fixed by experiment \cite{higgsmass}. Based on our choice of $\beta$ and experimental constraints \cite{2hdmreview}, we choose \begin{eqnarray} m_1=125\,\textrm{GeV}, m_2=300\,\textrm{GeV}, m_3=350\,\textrm{GeV},m_{\pm}=400\,\textrm{GeV}.\label{equ:masses} \end{eqnarray} \end{itemize} This leaves a 4-dimensional parameter space, spanned by $\alpha_{1}$, $\alpha_{2}$, $\theta$, and $\mu$. \subsubsection{Symmetries} \label{sec:symmetries} Symmetries in the Higgs potential $V$ help us to further simplify our calculations. Since $V$ is real, it follows that \begin{equation} \{ \lambda \} \to \{ \lambda \}^*,\qquad \Phi_i \xrightarrow{C} \Phi_i^*,\label{equ:complexConj} \end{equation} is a symmetry. This imposes the relation between the sets of parameters $\{\lambda\}$, \begin{eqnarray} \label{eq:alpha3symmetry} \{\lambda\}[\alpha_1,\alpha_2,\alpha_3]=\left(\{\lambda\}[\alpha_1,-\alpha_2,\pi-\alpha_3]\right)^*, \end{eqnarray} which is equivalent to the charge conjugation of the bosonic fields according to (\ref{equ:complexConj}). Therefore, the generated baryon asymmetry flips sign when complex conjugating the parameter set $\{\lambda\}$, i.e., \begin{eqnarray} \label{eq:alpha3symmetry2} &&N_{\rm cs}[\alpha_1,\alpha_2,\alpha_3]=-N_{\rm cs}[\alpha_1,-\alpha_2,\pi-\alpha_3],\nonumber\\ &&N_{\rm w}^{1,2}[\alpha_1,\alpha_2,\alpha_3]=-N_{\rm w}^{1,2}[\alpha_1,-\alpha_2,\pi-\alpha_3], \end{eqnarray} and so there is a redundancy between the upper and lower half-plane in $\alpha_1-\alpha_2$ space. Finally, the symmetry \begin{equation} \phi_1 \to e^{-i \theta} \phi_1,\qquad\lambda_5\to e^{-2i\theta}\lambda_5,\qquad\mu_{12}^2 \to e^{-i\theta} \mu_{12}^2. \label{sec:thetatrans} \end{equation} will also be very useful. Using this transformation, one can easily see that \begin{equation} \lambda_5[ \theta ] = e^{2i\theta} \lambda_5[ 0 ],\qquad\mu_{12}^2[ \theta ] = e^{i\theta} \mu_{12}^2[ 0 ].\label{equ:theta} \end{equation} Therefore, one can first find the parameter set $\{ \lambda \}[0]$ and then obtain $\{ \lambda \}[\theta]$ by the above transformation. The physical Higgs masses are unchanged under such a transformation. Since $\mu$ is invariant under (\ref{equ:theta}), we need only consider varying the potential in the 3-dimensional parameter space spanned by $(\alpha_1, \alpha_2, \mu)$, and we get the $\theta$-direction for free. The potential at different values of $\theta$ are equivalent, but with different field basis. However, the symmetry in (\ref{equ:theta}) is explicitly broken as soon as the scalar sector is coupled to fermions, or in our case the C-/P-violating term in (\ref{eq:CandP}) is included. Then different $\theta$ are physically distinct, as under the transformation (\ref{sec:thetatrans}), \begin{equation} \phi_1^\dagger\phi_2-\phi_2^\dagger\phi_1\rightarrow e^{i\theta}\phi_1^\dagger\phi_2-e^{-i\theta}\phi_2^\dagger\phi_1. \end{equation} \subsubsection{Basic constraints, maxima and saddle points} \label{sec:maxsad} For a given value of $\mu$, we now survey the whole $\alpha_1-\alpha_2$ plane. For each such pair, we accept/reject based on overall stability (potential is bounded from below), unitarity (tree-level Higgs-Higgs scattering amplitudes are smaller than unity), and whether the minimum found is a global minimum. Conditions for stability and unitarity are well-known and the interested reader is referred to \cite{2hdmreview} and references therein. For the requirement of the global minimum, we find all the other minima of the potential and establish that the chosen one is in fact the one with lowest potential energy. We also reject if there are no solutions for $\alpha_3$, and finally we reject if the potential has a minimum at $(\phi_1,\phi_2)=(0,0)$ (see below). The origin $(\phi_1,\phi_2)=(0,0)$ is always a stationary point of $V$. For each of the surviving pairs $(\alpha_1,\alpha_2)$, we compute the eigenvalues of the mass matrix at the origin (not to be confused with $M^2$ of (\ref{eq:M2}), the neutral Higgs sector mass matrix in the minimum), \begin{eqnarray} \mathcal{M}^2=-\frac{1}{2}\left(\begin{array}{cc} \mu_{11}^2&\mu_{12}^2\\ \mu_{12}^{2,*}&\mu_{22}^2 \end{array}\right). \end{eqnarray} If both eigenvalues are negative, both Higgs fields will experience a spinodal transition, and we name this parameter point a {\it maximum}. If only one eigenvalue is negative (and the other positive), only one field goes spinodal, and we name the parameter point a {\it saddle point}\footnote{Note that since both fields acquire expectation values, eventually also the second field must undergo symmetry breaking, but then as a result of the first field going through its spinodal transition.}. If both eigenvalues are positive, no spinodal instability occurs and we reject the point. In principle, such a minimum could lead to tunneling and bubble nucleation on the way to symmetry breaking, but this returns us to standard electroweak baryogenesis, which we do not consider here (but see also \cite{servant}). \begin{figure} \begin{center} \includegraphics[width=15cm]{alphadivmu.eps} \end{center} \caption{The allowed values of $\alpha_{1,2}$ after all constraints have been applied, for different values of $\mu$. Black dots are maxima, red dots are saddle points. Two superposed points refer to the two different allowed values of $\alpha_3$. We perform simulations at every second allowed point at $\mu=100\,\text{GeV}$.}\label{fig:alphadivmu} \end{figure} Fig.~ \ref{fig:alphadivmu} shows the $\alpha_1-\alpha_2$ plane for various values of $\mu$, where we have sampled points with a spacing of $\pi/40$. We have indicated maxima by fat black dots, and saddle points by smaller red dots. The rest of the parameter space has been discarded for one of the reasons explained above. Where a red and a black dot are superposed, this corresponds to the two values of $\alpha_3$, and that these give a maximum and a saddle point, respectively. The lines $\alpha_2=0,\pm\pi/2$ have zero CP-violation, and can therefore not provide baryogenesis. At zero $\mu$, no choice of $\alpha_{1,2}$ survives the constraints. For small, but non-zero $\mu$, the allowed region is close to the $\alpha_1$-axis. For $\mu=100\,$GeV, about a third of the off-axis points are maxima, the rest are saddle points. As $\mu$ is further increased, the allowed region spreads out to a band near $\alpha_1=0.5$ which reconnects around the circle in the $\alpha_2$-direction. A ``hole'' also opens up around the origin. At the largest $\mu$, all off-axis points are saddle points, and by $\mu=400\,$ GeV, no points survive. Interestingly, by far the most important constraint is that the minimum should be the global minimum. All but a few of the discarded points fail in this respect. We expect that a similar picture arises for other choices of $m_{1,2,3,\pm}$ with the allowed region shifted accordingly in $\mu-\alpha_{1}-\alpha_{2}$-space. We now turn to our numerical lattice simulations, where we have computed the baryon asymmetry for all the allowed parameter space for $\mu=100\,$GeV, top middle of Fig.~\ref{fig:alphadivmu}, but with a coarser spacing of $\pi/20$. \section{Numerical results: } \label{sec:fullnum} \begin{figure} \begin{center} \includegraphics[width=7cm]{Singletraj.eps} \includegraphics[width=7cm]{Average008_100.eps} \end{center} \caption{Left: The Chern-Simons number and Higgs winding numbers in a single initial field realization. The inset shows the Higgs field expectation values squared. Right: The Chern-Simons number and Higgs winding numbers averaged over the ensemble.}\label{fig:averageexample} \end{figure} The action (\ref{eq:action}) is discretized on a lattice and the classical equations of motion derived and solved numerically. Starting from a zero-temperature initial condition, we study the evolution of the system through the spinodal transition. Observables are averaged over a statistical ensemble of initial realizations, which is by hand C-, P- and CP-symmetric (for details, see \cite{paper01}). The baryon asymmetry is inferred from the anomaly equation (\ref{eq:anomalyeq}). In fact, because we are initially very far from equilibrium, Chern-Simons number is not a very clean observable, since it is in general non-integer and can exhibit large oscillations at intermediate times. Instead, we consider the Higgs winding number, which coincides with Chern-Simons number at late times, is integer throughout and settles much earlier into its late-time value. Since we have two Higgs fields, we also have two winding numbers, both of which will eventually match Chern-Simons number, and we identify the late-time value by the time at which the two agree, irrespective of the value of the Chern-Simons number. In \cite{paper01}, we studied the dependence of the baryon asymmetry on the strength of C-/P-violation. In the present paper, we fix $\delta_{C/P}=-21$ which is close enough to the linear regime that we can interpolate to smaller values \cite{paper01} ($\delta_{\rm C/P}=0$ gives zero asymmetry by construction). In this way, we can investigate the significance of CP violation by studying the dependence of the baryon asymmetry on $\alpha_{1,2,3}$ and $\theta$. Fig.~\ref{fig:averageexample} (left) shows the evolution of winding numbers and Chern-Simons number for one particular configuration for a particular choice of $(\alpha_1,\alpha_2,\alpha_3)=(0,\pi/10,1.138)$. Winding number has clearly settled, while Chern-Simons number is still catching up. In the inset, we show the Higgs field expectation values squared, of which one settles very rapidly, and one keeps oscillating for a long time, and with large amplitude. This is because the potential around the minimum is steep in one direction and shallow in the other. In Fig.~\ref{fig:averageexample} (right) we show the average winding number and Chern-Simons number, averaged over 100 sets of 4 conjugate configurations. Most of the high-frequency noise in the winding numbers has been averaged out, and the two nicely settle at a common value, quite early on in the evolution. By $vt=30$, symmetry breaking is complete. We also see that statistical errors are well under control at this size of ensemble. The average Chern-Simons number, however, has certainly not settled to its equilibrium value. Two effects are at work here: There is a net shift downwards, which is a transient non-equilibrium effect. We checked, by running for three times as long, that eventually the Chern-Simons number settles to the winding number value. The second effect is a large-amplitude oscillation, and is the result of the C-/P-violating driving force being given by the oscillating Higgs field vevs. We see from the figure that the oscillation has the same frequency as the Higgs field oscillations, but are shifted by a phase. This follows from considering the C-/P-violating term as \begin{eqnarray} S_{\rm C/P}\propto \delta_{\rm C/P}\,\textrm{Im}[\phi_1^\dagger\phi_2]\, \partial_t N_{\rm cs}, \end{eqnarray} which holds approximately for almost homogeneous Higgs fields. By partial integration, this term can be considered a time dependent driving force or chemical potential for Chern-Simons number, with magnitude $\propto \delta_{\rm C/P}\partial_t(\textrm{Im}[\phi_1^\dagger\phi_2])$. The reason why this second effect is not washed out by the averaging procedure is that all members of the ensemble experience (roughly) the same oscillation frequency and phase of the driving force, since the Higgs oscillation is almost universal, configuration by configuration. Therefore, although other configuration-specific effects average out to give a small baryon asymmetry, the driven oscillation is common to all configuration and survives the averaging process. At late times, the Higgs fields will also stop oscillating, and the driving force will disappear. But even at these early times, the coherent oscillation has no impact on the average winding numbers, which we therefore take as our measurement of the generated baryon asymmetry. \subsection{Symmetry under $C$ and $\lambda\rightarrow \lambda^*$} \label{sec:symC} \begin{figure} \begin{center} \includegraphics[width=7cm]{lambdaconjcheck.eps} \end{center} \caption{Higgs winding number for a parameter set $\{ \lambda \}$ and for $\{ \lambda \}^*$, with an overall flipped sign. Here, $(\alpha_1, \alpha_2, \mu, \theta)=(0, \frac{\pi}{10}, 100\,\text{GeV}, -1.27)$ and the results are averaged over an ensemble of $4\times25$ configurations. }\label{fig:Csym} \end{figure} In Fig.~\ref{fig:Csym}, we demonstrate explicitly that the symmetry (\ref{eq:alpha3symmetry2}) holds, by simply computing the asymmetry for a parameter set $\{\lambda\}$ and its complex conjugate, and then flipping the sign of the resulting asymmetry. We see that the agreement is very good (within statistical error bars). Hence we find the advertised redundancy between positive and negative values of $\alpha_2$. \subsection{Dependence on $\theta$} \label{sec:theta} \begin{figure} \begin{center} \includegraphics[width=7cm]{thetadep.eps} \end{center} \caption{The dependence of the winding number asymmetry on $\theta$ for $(\alpha_1, \alpha_2, \mu)=(0, \frac{\pi}{10}, 100\,\text{GeV})$ (black and red) and $(\frac{2\pi}{5}, \frac{\pi}{20}, 100\,\text{GeV})$ (green and blue).}\label{fig:theta} \end{figure} The transformation (\ref{sec:thetatrans}) allows us, from a given set of parameters $\{\lambda\}$, to generate a whole set of identical potentials, but where the minimum is rotated to $v_1 e^{-i\theta}$ for any value of $\theta$. We generate $\{\lambda\}$ at $\theta=0$ using the constraint $\textrm{Im}\,\mu_{12}^2=v_1v_2\textrm{Im}\lambda_5$, and from the point of view of CP-violation, all values of $\theta$ are equivalent. But once we couple to C-/P-violation, the potentials are distinct. In the vacuum, we have \begin{equation} \mathcal{L}_{\rm C/P}\propto \delta_{\rm C/P}2\,v_1\,v_2\sin(\theta)\textrm{ Tr}\,F_{\mu\nu}\tilde{F}^{\mu\nu}, \end{equation} and so were we in vacuum throughout the transition, there would be no asymmetry generated at $\theta=0$. And naively, one would expect the asymmetry to be proportional to $\sin\theta$. Fig.~\ref{fig:theta} shows the asymmetry in $N_{\rm w}^2$ as a function of $\theta$ for the parameter sets $\{\lambda\}(\alpha_1, \alpha_2, \mu)=(0, \frac{\pi}{10}, 100~~\text{GeV})$ (black dots) and $(\frac{2\pi}{5}, \frac{\pi}{20}, 100\,\text{GeV})$ (green dots). The arrows indicate the values of $\theta$ corresponding to real $\lambda_5$, the criterion we will use for most of our simulations below. We have fit with a form $A \sin(\theta+\delta\theta)$, and find beautiful agreement, but with a non-zero $\delta\theta=2.26$ (red line) and $\delta\theta=1.8$ (blue line). As a result, even at $\theta=0$, an asymmetry is generated during the transition where $\theta$ is different from its vacuum value. This is a result of the Higgs fields rolling down the potential in a spinodal transition, where both the length and the phase of the fields vary locally, until they finally settle near their vacuum values. The surprising result is perhaps that the simple $\sin\theta$ form is preserved, and that the out-of-equilibrium stage is encoded in the $\theta$-dependence as an overall shift of the phase, $\delta\theta$. But this also means that the asymmetry vanishes at $\theta=-\delta\theta$ (and $\theta=\pi-\delta\theta$), and that the overall sign of the asymmetry varies in this simple way with $\theta$, presumable for any set $\{\lambda\}$. We do not know of an obvious way of determining $\delta\theta$ apart from through the simulations. On the other hand, since we can parameterize the dependence through $A$ and $\delta\theta$, we only need simulations at two values of $\theta$, say $\theta=0$ and $\theta=\pi/2$. Then we simply have \begin{equation} \tan(\delta\theta)=\frac{N_{\rm w}(\theta=0)}{N_{\rm w}(\theta=\pi/2)},\qquad A=\frac{N_{\rm w}(\theta=0)}{\sin(\delta\theta)}, \end{equation} from which one can find $\delta\theta$ and then $A$. In the following, we have for each $\{\lambda\}[0]$ rotated to the value of $\theta$ that gives a real $\lambda_5$. This choice is arbitrary, and can as we have seen with a comparable amount of additional some numerical effort be extended to a complete $\theta$-dependence. \subsection{Dependence on $\alpha_{1,2}$} \label{sec:alphadep} \begin{figure} \begin{center} \includegraphics[width=7cm]{Averageall2.eps} \includegraphics[width=7cm]{Averageall1.eps} \end{center} \caption{The final baryon asymmetry as a function of $\alpha_1$ for $\alpha_2=\pi/20$ (left) and $\pi/10$ (right).}\label{fig:alpha12} \end{figure} Fig.~\ref{fig:alpha12} is the baryon asymmetry as a function of $\alpha_{1}$ at $\mu=100\,$GeV for $\alpha_2=\pi/20$ (left) and $\pi/10$ (right). We have used the conversion from winding number to baryon number \begin{eqnarray} \frac{n_B}{n_\gamma}=1.2\times 10^{-4}\left(\frac{V(0,0)-V(v_1,v_2)}{v^4}\right)^{-3/4}\times \langle N_{\rm w}^{1,2}\rangle. \end{eqnarray} This assumes that the total potential energy is distributed onto all the Standard Model degrees of freedom with masses less than $m_{\rm w}$, giving the reheating temperature and the photon number density $n_\gamma$. As discussed in the previous section, the $\alpha_2>0$ and $\alpha_2<0$ regions are related by Eq.~(\ref{eq:alpha3symmetry}) Since there are at most two allowed values of $\alpha_3$ at each grid point in the $(\alpha_1, \alpha_2)$ plane, we finally need to perform our numerical simulation using 39 sets of parameters. At $\alpha_2=\pi/20$ (left) we see a clear peak close to $\alpha_1=0$, which gradually decreases toward the edges of the allowed parameter region. For large $\alpha_1$ the results are roughly consistent with zero. The two values of $\alpha_3$ happen to give very similar results within errors. Note that connecting the largest/smallest $\alpha_3$-results by curves is an arbitrary choice to guide the eye. At $\alpha_2=\pi/10$ (right) we first of all observe that the two values of $\alpha_3$ do not agree as well, although they are still within a factor of two. The asymmetry is larger than for $\alpha_2=\pi/20$, and shows no sign of smoothly going to zero at the edge of the parameter range. This may be a result of the coarse resolution in $\alpha_1$. Again, the connecting curves are just to guide the eye. We see that the maximum value is again attained near $\alpha_1=0$, although now a peak structure is less clear. We should also keep in mind the $\theta$-dependence, and that the results shown here could be at any point in the period of the $\sin\theta$ behaviour. The small remaining parameter range at large $\alpha_2$ (see Fig.~\ref{fig:alphadivmu}) gives baryon asymmetries of roughly the same size, and are by no means suppressed compared to small $\alpha_2$. We should also mention that we checked that the magnitude of the asymmetry is not in a simple way correlated with the determinant or eigenvalues of the mass matrix $\mathcal{M}^2$ at the origin, and in particular whether we start at a maximum or a saddle point. There is also no simple correlation with the phase of $\mu_{12}^2$ or of $v_1$. We did, however find a weak correlation in the combined $\textrm{Im}(v_1)$-$(V(0,0)-V(v_1,v_2))$-plane, suggesting that a deep potential drop and large CP-violation gives a large baryon asymmetry. This is perhaps not unexpected, but is surprisingly difficult to confirm. Clearly, the complicated non-linear dynamics does not allow for such simple conclusions about the generated asymmetry. \section{Conclusion and outlook} \label{sec:conclusion} We have outlined a practical parametrization of the 4-dimensional parameter space in the 2HDM resulting from fixing masses and vevs. We have seen that imposing a number of general consistency criteria, a finite region in $\mu-\alpha_1-\alpha_2$ survives, and this can be extended to the $\theta$-direction by a simple phase change transformation. The parameter space is a hyper-cylinder with $\mu>0$ and three angles $\alpha_1$, $\alpha_2$ and $\theta$, with the additional redundancy that $\alpha_2>0$ and $\alpha_2<0$ are connected. On the other hand each set of $\alpha_{1,2}$ has up to two solutions for $\alpha_3$. When the bosonic sector is coupled to fermions, different values of $\theta$ are physically distinct, but the asymmetry seems to follow a form $A\sin(\theta+\delta\theta)$. $\delta\theta$ is a priori unknown, but can be found numerically by using two different values of $\theta$, say spaced by $\pi/2$. Finally, it seems that the extent in $\mu$ is finite and determined by the overall scale of the fixed Higgs masses. The maximal asymmetry we found is at the point $(\alpha_1,\alpha_2,\alpha_3)=(\pi/20,\pi/10,1.333)$, and is \begin{eqnarray} \left(\frac{n_B}{n_\gamma}\right)_{\rm max}=-1.1\times 10^{-5}\times \delta_{\rm C/P}, \end{eqnarray} so that in order to reproduce that observed asymmetry of $\sim 6\times 10^{-10}$, we require \begin{eqnarray} \delta_{\rm C/P}\simeq (5\textrm{ to }6) \times 10^{-5}. \end{eqnarray} We note that the maximal asymmetry is a factor 3 or 4 larger than what we found in \cite{paper01} at one particular parameter point, which therefore was not a particularly unique case. Also, there are parts of the allowed parameter space that give vanishing asymmetry. Computing $\delta_{\rm C/P}$ from first principles could therefore potentially rule out regions of 2HDM parameter space, under the assumption that baryogenesis originates from a cold spinodal transition involving two Higgs field. Although a similar programme could be attempted for other scenarios of baryogenesis, these are less amenable to a direct, quantitative computation. Leptogenesis is a multi-stage process, generation of lepton asymmetry, thermalization, sphaleron processes, freeze-out. And ``Hot'' electroweak baryogenesis involves the nucleation of bubbles, their interaction with the plasma and again sphaleron processes. Cold electroweak baryogenesis offers a practicable testing groud for this kind of parameter scans. Given the numerical effort involved in the present work (of order $10^5$ CPU hours on a standard linux cluster), it is difficult to scan through the currently allowed parameter space, including the remaining Higgs masses and $\tan \beta$. But a complete sweep of the 4-dimensional parameter space can be done with about a factor of 10-100 more computing power, which is easily within reach of current supercomputers. And hopefully, the coming years of LHC-experiments at the electroweak energy scale will constrain the viable range of masses and vevs, or even discover additional scalar particles. When this happens, it would be natural to revisit the scenario considered here, and use the baryon asymmetry to narrow down the range of experimentally less accessible parameters. \acknowledgments We would like to thank Aleksi Vuorinen, Tomas Brauner and Olli Taanila for helpful discussions. B.W. was supported by the Humboldt foundation through its Sofja Kovalevskaja program. A.T. was supported by the Carlsberg Foundation and the Villum Kann Rasmussen Foundation.
\section{Tensor Product Structures} Operationally, entanglement is detected as the coherence between non-local observables, but how do we define local? The classic scenario presents Alice and Bob, probing spin states in locations with space-like separation. By comparing correlated expectations values for two pairs of non-orthogonal measurements, phase correlation among the expectation values, i.e.\ coherence, is revealed. In this case, the notion of local subsystems coincides with the standard geometric notion of local. More generally, one can define local by the subsystems themselves. For example, though two particles in a trap are strongly interacting,the Hilbert space of the system is constructed as the tensor product of each particle's Hilbert space. The particles are certainly not separated in space-time, but they may or may not have entangled wave functions by this `natural' notion of subsystem locality and separability. How far can we take the subsystem definition of locality? Zanardi's Theorem~\cite{zanardi}, gives the mathematical and physical requirements for a partition of the algebra of observables $\mathcal{A}$ into subalgebras $\{\mathcal{A}_i\}$ to induce a tensor product structure (TPS) of virtual subsystems\footnote{Defining entanglement based on observable-induced virtual subsystems is not the only extension of separability and entanglement possible. The term `generalized entanglement' refers to an observable-induced definition of entanglement that does not imply a partition in to subsystems or a tensor product structure~\cite{viola}.}. Given a state space $\Phi\subset\mathcal{H}$, then the Theorem says a collection of subalgebras $\{\mathcal{A}_1,\mathcal{A}_2,\ldots \}$ of the total operator algebra $\mathcal{A}$ acting on $\mathcal{H}$ will induce a tensor product structure $\mathcal{H}=\bigotimes_i \mathcal{H}_i$ if they satisfy the following criteria: subsystem independence, completeness, and local accessibility. The first two conditions are mathematical criteria. Subsystem independence means that all the subalgebras commute and completeness means that the subalgebras generate the total algebra. However, local accessibility is a physical criteria: each subalgebra must correspond to a set of controllable observables. Zanardi's Theorem is currently proven only for Hilbert spaces realized by $\mathbb{C}^d$ with $d$ finite. The total algebra of observables is realized as a subset of $\mathcal{A}\cong\mathbb{M}^d$, the complex $d\times d$ matrices. Assume $d$ can be factored into positive integers as $d=\prod_i k_i$, then the subalgebras $\mathcal{A}_1 \cong \mathbb{M}^{k_1}\otimes\mathbb{I}_{k_2}\otimes\mathbb{I}_{k_3}\cdots$, $\mathcal{A}_2 \cong \mathbb{I}_{k_1}\otimes\mathbb{M}^{k_2}\otimes\mathbb{I}_{k_3}\cdots$, etc.\ induce the TPS $\mathcal{H} \cong \mathcal{H}_1 \otimes \mathcal{H}_2 \otimes \mathcal{H}_3 \cdots$. This construction is not unique, so for every factorization of $d$, there is an equivalence class of TPSs. Even in the simplest non-trivial case where $d=4$, alternate partitions of the observable algebra can lead to alternate evaluations of entanglement for a given state. The Bell states $|\Phi^\pm\rangle$ and $|\Psi^\pm\rangle$ are maximally entangled with respect to commuting observable subalgebras of Alice $\mathcal{A}\cong\mathrm{span}\{{\sigma}^A_i\otimes\mathbb{I}^B\}$ and Bob $\mathcal{B}\cong\mathrm{span}\{\mathbb{I}^A\otimes\sigma_i^B\}$. These induce a TPS $\mathcal{H}=\mathcal{H}_A\otimes\mathcal{H}_B$. However, any unitary operator $U$ that cannot be factored into local unitaries can be used to make new subalgebras $\mathcal{P}=U\mathcal{A}U^\dag$ and $\mathcal{Q}=U\mathcal{B}U^\dag$ that induce a new TPS $\mathcal{H}=\mathcal{H}_P\otimes\mathcal{H}_Q$. For a suitable choice of $U$, the $PQ$-type entanglement for a pure state using any measure can be anything from maximal to none~\cite{maxnone}. For pure states, perhaps the furthest expression of this TPS-relativity is the Tailored Observables Theorem~\cite{harshman_observables_2011}: for any Hilbert space $\mathbb{C}^d$ one can construct tailored subalgebras of observables that induce a TPS from a finite basis of operators such that any known pure state can have any entanglement that is possible for any prime factorization of $d$. The proof in \cite{harshman_observables_2011} is by construction: for a known $|\psi\rangle\in\mathbb{C}^d$ and chosen factorization of $d$, generators for the inducing subalgebras can be constructed in finite steps. Can these results for entanglement relativity be extended beyond pure states? The totally mixed state is invariant under this kind of relativity of observables because $U$ commutes with the mixed state $\rho=\mathbb{I}^d$. Mixed states are less coherent than pure states, as quantified by $\mathrm{Tr}(\rho^2)$, and therefore have less flexibility to capture entanglement by adjusting the TPS. For finite-dimensional Hilbert space realizations, the procedure and scope of entanglement relativity has only been explored in detail for $d=4$~\cite{thirring}. There are a few results for entanglement relativity in Gaussian states that can be inferred from the literature. For example, there is always some symplectic transformation $S$ that diagonalizes the covariance matrix. This phase space transformation can be represented as a unitary operator $U_S$ in the Hilbert space of the system. In the case of a two-mode Gaussian state, for example, the operator $U_S$ can be used to transform the canonical Heisenberg algebras $\mathcal{A}_i$ generated by $\{X_i,P_i\}$ for each mode into commuting subalgebras $U_S\mathcal{A}_i U_S^\dag$. These new subalgebras induce a TPS with respect to which the known Gaussian state is separable. This argument holds true for pure and mixed Gaussian states, and provides additional confirmation that Gaussian states are ``the most classical'' of quantum states. Beyond Gaussian states, the limits of observable-induced entanglement relativity in infinite-dimensional Hilbert space realizations is an open mathematical question. How do you classify the entanglement possibilities and what observables induce them? Given a pure or mixed state, how can one mathematically construct TPSs that minimize and maximize entanglement? Can observables be constructed that would extract some kind of entanglement from simple, physically-accessible systems? \section{Symmetries of the Two Body System} Observable subalgebras constructed to tailor the entanglement of a quantum state can be shown to exist using Zanardi's notion of induced TPSs, but these observables may not satisfy Zanardi's third condition of local accessibility. For systems like linear quantum optics, correlated detector measurements may allow the experimentalist to reconstruct expectation values for non-local observables and realize tailored observables in the laboratory, but for interacting particle systems we cannot access the ``experimental knobs'' that would allow us to tune our observable subalgebras. The symmetries of space and time, and the properties of the interaction and environment, determine which observables are most accessible and useful. Consider a typical quantum two-body system, such as a proton and electron in free space, interacting via the Coulomb potential. All electron observables $\mathcal{A}_{\rm e}$ are generated by a set of free particle operators, e.g.\ $\{{\bf X}^{\rm e},{\bf P}^{\rm e},{\bf S}^{\rm e}\}$. A similar set generates $\mathcal{A}_{\rm p}$ for the proton. The vector operators for position ${\bf X}$, momentum ${\bf P}$ and intrinsic spin ${\bf S}$ are themselves constructed from the Lie algebra of Galilean spacetime transformations. These are natural observables when the electron and proton are far apart and weakly-interacting, but for bound states these observables are less accessible. The total algebra of observables $\mathcal{A}$ is the direct sum $\mathcal{A}_{\rm e}\oplus\mathcal{A}_{\rm p}$ and the tensor product structure induced is $\mathcal{H}=\mathcal{H}_{\rm e}\otimes\mathcal{H}_{\rm p}$. Because of this construction, Galilean transformations are represented as local unitary operators with respect to this TPS and therefore interparticle entanglement is a Galilean invariant for pure states and mixed states. In elastic scattering, one generally assumes the asymptotic in-state has no interparticle entanglement, and therefore scattering can only increase it. One can show in a variety of model systems that almost every scattering interaction leads to an increase in entanglement for almost every unentangled in-state~\cite{scattering}. For bounds states, interparticle entanglement is also the norm, even for the simple system of coupled harmonic oscillators~\cite{osc}. Each particle Hilbert space is realized as square-integrable/square-summable functions on the spectra of a complete set of observables, for example the choice $\{{\bf X}^{\rm e},{S}_z^{\rm e}\}$ gives us the realization $\mathcal{H}_{\rm e}=\mathcal{H}^{\bf x}_{\rm e}\otimes\mathcal{H}^s_{\rm e} = \mathrm{L}^2(\mathbb{R}^3)\otimes\mathbb{C}^2$. More generally, the subalgebras $\mathcal{A}^{\bf x}_{\rm e}$ and $\mathcal{A}^{\rm s}_{\rm e}$ generated by $\{{\bf X}^{\rm e},{\bf P}^{\rm e}\}$ and $\{{\bf S}^{\rm e}\}$ are complete and commuting, and they can be used to defined the tensor product structure for intraparticle entanglement between motional and spin degrees of freedom for a single particle. For the electron, the maximum entanglement with respect to this TPS is the same as for a TPS of two qubits~\cite{continuous}, and it is also a Galilean invariant~\cite{osid}. The Hamiltonian $H$ is an operator in $\mathcal{A}$ that cannot be expressed as the sum of an electron operator in $\mathcal{A}_{\rm e}$ and and a proton operator in $\mathcal{A}_{\rm p}$. However, we know that there is a special symplectic transformation $S$ compatible with the overall Galilean symmetry that separates a Hamiltonian with a central interaction. In phase space, the transformation $S$ is a linear map that takes the canonical coordinates $({\bf x}^{\rm e},{\bf x}^{\rm p},{\bf p}^{\rm e},{\bf p}^{\rm p})$ into the center-of-mass and relative coordinates $({\bf x}^{\rm c},{\bf x}^{\rm r},{\bf p}^{\rm c},{\bf p}^{\rm r})$. The unitary representation of this transformation $U_S$ leaves invariant the total algebra $\mathcal{A}=U_S \mathcal{A} U_S^\dag$ and each particle's spin subalgebra $\mathcal{A}^{\rm s}_{\rm e}$ and $\mathcal{A}^{\rm s}_{\rm p}$. The representation $U_S$ linearly transforms the generators of the motional subalgebras $\mathcal{A}^{\bf x}_{\rm e}$ and $\mathcal{A}^{\bf x}_{\rm p}$ into the generators of the center-of mass $\mathcal{A}^{\bf x}_{\rm c}$ and relative $\mathcal{A}^{\bf x}_{\rm r}$ motional observables. The decomposition of $\mathcal{A}$ into $\mathcal{A}^{\bf x}_{\rm c}\oplus\mathcal{A}^{\bf x}_{\rm r}\oplus \mathcal{A}^{\rm s}_{\rm e}\oplus\mathcal{A}^{\rm s}_{\rm p}$ provides another Galilean invariant TPS. The transformed central-force two-body Hamiltonian $U_S {H} U_S^\dag$ decomposes into center-of-mass Hamiltonian ${H}^{\rm c}\in\mathcal{A}^{\bf x}_{\rm c}$, plus the relative Hamiltonian ${H}^{\rm r}\in\mathcal{A}^{\bf x}_{\rm r}$. This separability is a quantum manifestation of classical two-body solvability; Galilean symmetry applied to phase space provides enough independent Casimir invariants to integrate the motion. For the center-of-mass (or external) Hilbert space $\mathcal{H}_{\rm ext}=\mathcal{H}^{\bf x}_{\rm c}$, a complete set of commuting observables is $\{{\bf X}^{\rm c}\}$ and wave functions are square-integrable functions on their spectrum $\mathbb{R}^3$. The relative Hamiltonian $\hat{H}^{\rm r}$ for a central, spin-independent interaction acts only on $\mathcal{H}_{\rm rel}=\mathcal{H}^{\bf x}_{\rm r}$. One could realize this Hilbert space as functions on the spectrum of the complete commuting set $\{{\bf X}_{\rm r}\}$, but these operators do not commute with the relative Hamiltonian. A typical set would be $\{{H}_{\rm r},\hat{L}^2,\hat{L}_z\}$, and the Hilbert space spectrum of ${H}^{\rm r}$ decomposes into a discrete, accumulating set of bound states and a continuum of scattering states. More generally, a spin-dependent non-central Hamiltonian would be an operator in the internal subalgebra $\mathcal{A}_{\rm int}=\mathcal{A}^{\bf x}_{\rm r}\oplus\mathcal{A}^{\rm s}_{\rm e}\oplus\mathcal{A}^{\rm s}_{\rm p}$. The induced TPS $\mathcal{H}=\mathcal{H}_{\rm ext}\otimes\mathcal{H}_{\rm int}$ is not just Galilean invariant, but it is dynamically invariant because the Hamiltonian exponentiates to a time evolution operator that is local with respect to the internal-external subalgebras~\cite{prl}. For a hydrogen atom, the center-of-mass motion will remain unentangled with the internal atomic states as long as there are no external fields that couple them. For example, a non-uniform external (classical) electrostatic or magnetostatic field will cause a dipole force that depends on the relative state, possibly leading leading to changes in the internal-external entanglement. For some external fields and some states, e.g.\ harmonically-trapped, equal mass particles, the internal-external entanglement will also be time-invariant~\cite{osc}. This example shows that the entanglement with respect to TPSs induced by certain observable subalgebras can be more useful for describing the quantum two-body system than others. Depending on the space-time symmetries and the particular system dynamics, different types of entanglement are invariant. In isolated hydrogen atoms, we expect lower internal-external entanglement than interparticle entanglement for typical states. This can be extended to harmonic oscillators, whose Gaussian ground states are unentangled in the mode coordinates, but highly entangled in the particle coordinates. In other words, using the `right' observables minimizes the entanglement. They shift the coherence and quantum correlations to `local' observables. Whether this notion can be used to find efficient schemes for approximating non-integrable systems, for example the three-body system, seems like an interesting topic for future research. \bibliographystyle{aipproc}
\section{Introduction} This paper aims to study Riesz kernels and pseudodifferential equations a\-ttached to quadratic forms over $p$-adic fields. The Riesz kernels are naturally connected with several types of (pseudo) differential equations in the Archimedean setting, see e.g. \cite{dRham}, \cite{K-V}, \cite{Rubin}, \cite{Riesz}, and non Archimedean one, see e.g. \cite{A-K-S}, \cite{Koch}, \cite{R-Zu}, \cite{R-Zu1}, \cite{Taibleson}, \cite{V-V-Z}. In particular, in the non Archimedean setting, Riesz kernels attach to `polynomials of degree one' has been used to solve pseudodifferential equations \cite{A-K-S}, \cite{Koch}, \cite{V-V-Z}. Our initial motivation was to extend \ these results to the case of polynomials of higher degree to obtain $p$-adic analogs of the results of \cite{dRham}, \cite{Riesz}. To present our results consider the diagonal quadratic form $f\left( \xi\right) =a_{1}\xi_{1}^{2 +\ldots+a_{n}\xi_{n}^{2}$, the local zeta function attached to $f$ is the distribution defined b \[ \left( \left\vert f\right\vert _{p}^{s},\phi\right) {\displaystyle\int\limits_{\mathbb{Q}_{p}^{n}\smallsetminus f^{-1}\left( 0\right) }} \left\vert f\left( \xi\right) \right\vert _{p}^{s}\phi\left( \xi\right) d^{n}\xi\text{, }\operatorname{Re}\left( s\right) >0. \] These distributions, called local zeta functions, were introduced in the 50's by I. Gel'fand and A. Weil, see \cite{G-S}, \cite{Igusa}. A Riesz kernel is a local zeta function multiplied by a suitable gamma factor. In the cases in which $n=2$, $4$ and the quadratic form is elliptic,\ we show that the Riesz kernels, considered as distributions on certain $p$-adic Lizorkin spaces, form an Abelian group under the operation of convolution, see Theorem \ref{mainA} and Remark \ref{nota_dim_2}. The proof of this fact depends on a theorem of Rallis-Schiffmann that asserts that the distributions of type $\left\vert f\right\vert _{p}^{s}$ satisfy certain functional equations, see \cite{R-S}, also \cite{Igusa3}, \cite{Sato-Shi}, \cite{FSato1}, \cite{FSato2}. In order to use this result we compute all the gamma factors that appear in the functional equation for $p$-adic quadratic forms of type $a_{1}\xi_{1}^{2}+a_{2}\xi _{2}^{2}+\cdots+a_{n}\xi_{n}^{2}$, see Theorem \ref{funcional}. As consequence, we obtain \ fundamental solutions for certain\ pseudodifferential equations, see Theorem \ref{mainB} an Remark \ref{nota_ultima}. The fundamental solutions presented here are `classical solutions', see Definition \ref{Classical_sol}, while those present in \cite{Koch} and \cite{Z-G1 -\cite{Z-G5}\ are `weak solutions'. We also obtain the existence of a pseudodifferential operator $\boldsymbol{f}\left( \partial,1\right) $, acting on a space of Lizorkin distributions, and a gamma factor $A(s)$ such that $\boldsymbol{f}\left( \partial,1\right) \left\vert f\right\vert _{p}^{s+1}=A(s)\left\vert f\right\vert _{p}^{s}$, where $f$ is an elliptic quadratic form of dimension $2$ or $4$, see Theorem \ref{mainB} and Remark \ref{nota_ultima}. This is a non Archimedean pseudodifferential (particular) version of a celebrated result of Sato-Bernstein, see \cite{Sato-Shi}, \cite{Igusa}, \cite[Section 6.1.2]{Z-G5}. Thus, a natural problem is to study the existence of pseudodifferential Sato-Bernstein operators, and the corresponding functions, in the setting \ $p$-adic prehomogeneous vector spaces, this problem was posed in \cite[Section 6.1.2]{Z-G5}. Finally, the results obtained here can be applied to study other types of pseudodifferential equations, these results will appear in a separate article elsewhere. \textbf{Acknowledgement}. The authors want to thank to Professor Fumihiro Sato for his kind assistance on the functional equation for the local zeta function attached to a quadratic form. In particular, we are very grateful to him for allowing us to use some of the ideas of his unpublished manuscript \cite{FSato3}. The second author wants to thank to Professor Sergii Torba for several useful comments and discussions about this article. The authors also want to thank to the referee for his/her careful reading of the article. \section{\label{Section1}Preliminaries} In this section we fix the notation and collect some basic results on $p$-adic analysis that we will use through the article. For a detailed exposition on $p$-adic analysis the reader may consult \cite{A-K-S}, \cite{Taibleson}, \cite{V-V-Z}. \subsection{The field of $p$-adic numbers} Along this article $p$ will denote a prime number different from $2$. The field of $p-$adic numbers $\mathbb{Q}_{p}$ is defined as the completion of the field of rational numbers $\mathbb{Q}$ with respect to the $p-$adic norm $|\cdot|_{p}$, which is defined as \[ |x|_{p} \begin{cases} 0 & \text{if }x=0\\ p^{-\gamma} & \text{if }x=p^{\gamma}\dfrac{a}{b}, \end{cases} \] where $a$ and $b$ are integers coprime with $p$. The integer $\gamma:=ord(x)$, with $ord(0):=+\infty$, is called the\textit{ }$p-$\textit{adic order of} $x$. We extend the $p-$adic norm to $\mathbb{Q}_{p}^{n}$ by takin \[ ||x||_{p}:=\max_{1\leq i\leq n}|x_{i}|_{p},\qquad\text{for }x=(x_{1 ,\dots,x_{n})\in\mathbb{Q}_{p}^{n}. \] We define $ord(x)=\min_{1\leq i\leq n}\{ord(x_{i})\}$, then $||x||_{p =p^{-\text{ord}(x)}$. Any $p-$adic number $x\neq0$ has a unique expansion $x=p^{ord(x)}\sum_{j=0}^{\infty}x_{i}p^{j}$, where $x_{j}\in\{0,1,2,\dots ,p-1\}$ and $x_{0}\neq0$. By using this expansion, we define \textit{the fractional part of }$x\in\mathbb{Q}_{p}$, denoted $\{x\}_{p}$, as the rational number \[ \{x\}_{p} \begin{cases} 0 & \text{if }x=0\text{ or }ord(x)\geq0\\ p^{\text{ord}(x)}\sum_{j=0}^{-ord(x)-1}x_{j}p^{j} & \text{if }ord(x)<0. \end{cases} \] For $\gamma\in\mathbb{Z}$, denote by $B_{\gamma}^{n}(a)=\{x\in\mathbb{Q _{p}^{n}:||x-a||_{p}\leq p^{\gamma}\}$ \textit{the ball of radius }$p^{\gamma }$ \textit{with center at} $a=(a_{1},\dots,a_{n})\in\mathbb{Q}_{p}^{n}$, and take $B_{\gamma}^{n}(0):=B_{\gamma}^{n}$. Note that $B_{\gamma}^{n (a)=B_{\gamma}(a_{1})\times\cdots\times B_{\gamma}(a_{n})$, where $B_{\gamma }(a_{i}):=\{x\in\mathbb{Q}_{p}:|x_{i}-a_{i}|_{p}\leq p^{\gamma}\}$ is the one-dimensional ball of radius $p^{\gamma}$ with center at $a_{i}\in \mathbb{Q}_{p}$. The ball $B_{0}^{n}(0)$ is equals the product of $n$ copies of $B_{0}(0):=\mathbb{Z}_{p}$, \textit{the ring of }$p-$\textit{adic integers}. \subsection{The Bruhat-Schwartz space} A complex-valued function $\varphi$ defined on $\mathbb{Q}_{p}^{n}$ is \textit{called locally constant} if for any $x\in\mathbb{Q}_{p}^{n}$ there exist an integer $l(x)\in\mathbb{Z}$ such tha \begin{equation} \varphi(x+x^{\prime})=\varphi(x)\text{ for }x^{\prime}\in B_{l(x)}^{n}. \label{local_constancy \end{equation} A function $\varphi:\mathbb{Q}_{p}^{n}\rightarrow\mathbb{C}$ is called a \textit{Bruhat-Schwartz function (or a test function)} if it is locally constant with compact support. The $\mathbb{C}$-vector space of Bruhat-Schwartz functions is denoted by $\mathbf{S}(\mathbb{Q}_{p}^{n})$. For $\varphi\in\mathbf{S}(\mathbb{Q}_{p}^{n})$, the largest of such number $l=l(\varphi)$ satisfying (\ref{local_constancy}) is called \textit{the exponent of local constancy of} $\varphi$. Let $\mathbf{S}^{\prime}(\mathbb{Q}_{p}^{n})$ denote the set of all functionals (distributions) on $\mathbf{S}(\mathbb{Q}_{p}^{n})$. All functionals on $\mathbf{S}(\mathbb{Q}_{p}^{n})$ are continuous. Set $\chi(y)=\exp(2\pi i\{y\}_{p})$ for $y\in\mathbb{Q}_{p}$. The map $\chi(\cdot)$ is an additive character on $\mathbb{Q}_{p}$, i.e. a continuos map from $\mathbb{Q}_{p}$ into $S$ (the unit circle) satisfying $\chi (y_{0}+y_{1})=\chi(y_{0})\chi(y_{1})$, $y_{0},y_{1}\in\mathbb{Q}_{p}$. Given $\xi=(\xi_{1},\dots,\xi_{n})$ and $x=(x_{1},\dots,x_{n})\in \mathbb{Q}_{p}^{n}$, we set $\xi\cdot x:=\sum_{j=1}^{n}\xi_{j}x_{j}$. The Fourier transform of $\varphi\in\mathbf{S}(\mathbb{Q}_{p}^{n})$ is defined as \[ (\mathcal{F}\varphi)(\xi)=\int_{\mathbb{Q}_{p}^{n}}\chi(-\xi\cdot x)\varphi(\xi)d^{n}x\quad\text{for }\xi\in\mathbb{Q}_{p}^{n}, \] where $d^{n}x$ is the Haar measure on $\mathbb{Q}_{p}^{n}$ normalized by the condition $vol(B_{0}^{n})=1$. The Fourier transform is a linear isomorphism from $\mathbf{S}(\mathbb{Q}_{p}^{n})$ onto itself satisfying $(\mathcal{F (\mathcal{F}\varphi))(\xi)=\varphi(-\xi)$. We will also use the notation $\mathcal{F}_{x\rightarrow\xi}\varphi$ and $\widehat{\varphi}$\ for the Fourier transform of $\varphi$. \subsection{Operations on Distributions} Let $\Omega$\ denote the characteristic function of the interval $\left[ 0,1\right] $. Then $\Delta_{k}\left( x\right) :=\Omega\left( p^{-k}\left\Vert x\right\Vert _{p}\right) $ is the characteristic function of the ball $B_{k}^{n}\left( 0\right) $. \subsubsection{Convolution} Given $f,g\in\mathbf{S}^{\prime}\left( \mathbb{Q}_{p}^{n}\right) $, their convolution $f\ast g$ is defined b \[ \left\langle f\ast g,\varphi\right\rangle =\lim_{k\rightarrow+\infty }\left\langle f\left( y\right) \times g\left( x\right) ,\Delta_{k}\left( x\right) \varphi\left( x+y\right) \right\rangle \] if the limit exists for all $\varphi\in\mathbf{S}\left( \mathbb{Q}_{p ^{n}\right) $. We recall that if $f\ast g$ exists, then $g\ast f$ exists and $f\ast g=g\ast f$, see e.g. \cite[Section VII.1]{V-V-Z}. In the case in which $g=\psi\in\mathbf{S}\left( \mathbb{Q}_{p}^{n}\right) $, \[ f\ast\psi\left( x\right) =\left\langle f\left( y\right) ,\psi\left( x-y\right) \right\rangle , \] see e.g. \cite[Section VII.1]{V-V-Z}. \subsubsection{Fourier transform} The Fourier transform $\mathcal{F}\left[ f\right] $ of a distribution $f\in\mathbf{S}^{\prime}\left( \mathbb{Q}_{p}^{n}\right) $ is defined b \[ \left\langle \mathcal{F}\left[ f\right] ,\varphi\right\rangle =\left\langle f,\mathcal{F}\left[ \varphi\right] \right\rangle \text{ for all }\varphi \in\mathbf{S}\left( \mathbb{Q}_{p}^{n}\right) \text{. \] The Fourier transform $f\rightarrow\mathcal{F}\left[ f\right] $ is a linear isomorphism from $\mathbf{S}^{\prime}\left( \mathbb{Q}_{p}^{n}\right) $\ onto $\mathbf{S}^{\prime}\left( \mathbb{Q}_{p}^{n}\right) $. Furthermore, $f=\mathcal{F}\left[ \mathcal{F}\left[ f\right] \left( -\xi\right) \right] $. \subsubsection{Multiplication} Set $\delta_{k}\left( x\right) :=p^{nk}\Omega\left( p^{k}\left\Vert x\right\Vert _{p}\right) $ for $k\in\mathbb{N}$. Given $f,g\in\mathbf{S ^{\prime}\left( \mathbb{Q}_{p}^{n}\right) $, their product $f\cdot g$ is defined b \[ \left\langle f\cdot g,\varphi\right\rangle =\lim_{k\rightarrow+\infty }\left\langle g,\left( f\ast\delta_{k}\right) \varphi\right\rangle \] if the limit exists for all $\varphi\in\mathbf{S}\left( \mathbb{Q}_{p ^{n}\right) $. We recall that \ the existence of the product $f\cdot g$ is equivalent \ to the existence of $\mathcal{F}\left[ f\right] \ast \mathcal{F}\left[ g\right] $. In addition, $\mathcal{F}\left[ f\cdot g\right] =\mathcal{F}\left[ f\right] \ast\mathcal{F}\left[ g\right] $ and $\mathcal{F}\left[ f\ast g\right] =\mathcal{F}\left[ f\right] \cdot\mathcal{F}\left[ g\right] $, see e.g. \cite[Section VII.5]{V-V-Z}. The following result will be used later on. \begin{lemma} [{\cite[Section VII.5]{V-V-Z}.}]\label{producdistributions}Let $f,g$ functions in $L_{loc}^{1}$ for which the functio \ {\displaystyle\int\limits_{\mathbb{Q}_{p}^{n}}} g\left( x\right) \varphi\left( x\right) f\left( x-\xi\right) d^{n}x\text{, \] is continuos at $\xi=0\in\mathbb{Q}_{p}^{n}$, for any\ $\varphi\in \mathbf{S}\left( \mathbb{Q}_{p}^{n}\right) $. Then the product $f\cdot g$ is in $\mathbf{S}^{\prime}\left( \mathbb{Q}_{p}^{n}\right) $ and the distribution is induced by the pointwise product $f\left( x\right) g\left( x\right) $. \end{lemma} \subsection{The Hilbert Symbol} The Hilbert symbol $(a,b)_{p}$, $a,b\in\mathbb{Q}_{p}^{\times}$, is defined b \[ (a,b)_{p}=\left\{ \begin{array} [c]{ll 1 & \text{if $ax^{2}+by^{2}-z^{2}=0$ has a solution }(x,y,z)\neq\left( 0,0,0\right) \text{ in $\mathbb{Q}_{p}^{3}$}\\ -1 & \text{otherwise. \end{array} \right. \] The Hilbert symbol possesses the following properties (see e.g. Theorem 3.3.1 \cite{Ki}) \begin{equation} (a,b)_{p}=(b,a)_{p}\text{ and }(a,c^{2})_{p}=1\text{, for }a,b,c\in \mathbb{Q}_{p}^{\times}; \label{HS_1 \end{equation} \begin{equation} (ab,c)_{p}=(a,c)_{p}(b,c)_{p}\text{, for }a,b,c\in\mathbb{Q}_{p}^{\times}; \label{HS_2 \end{equation} \begin{equation \begin{cases} (a,b)_{p}=1 & \text{for $a,b\in\mathbb{Z}_{p}^{\times}$}\\ (a,p)_{p}=\left( \dfrac{a_{0}}{p}\right) & \text{for $a\in\mathbb{Z _{p}^{\times},$ \end{cases} \label{HS_3 \end{equation} where $a_{0}\in\mathbb{Z}$, with $a\equiv a_{0}\operatorname{mod \mathbb{Z}_{p}$, and $\left( \dfrac{a_{0}}{p}\right) $ is the Legendre symbol. Along this article $\left[ \mathbb{Q}_{p}^{\times}\right] ^{2}$ denotes the subgroup of squares of $\mathbb{Q}_{p}^{\times}$. We recall that $\mathbb{Q}_{p}^{\times}/\left[ \mathbb{Q}_{p}^{\times}\right] ^{2}$ is a finite group with four elements. We fix $\left\{ 1,\epsilon,p,\epsilon p\right\} $ to be a set of representatives, here $\epsilon$\ is unit which is not square. It is clear that $\left( a,b\right) _{p}$ does not change when $a$ and $b$ are multiplied by squares, thus the Hilbert symbol gives rise a map from $\mathbb{Q}_{p}^{\times}/\left[ \mathbb{Q}_{p}^{\times}\right] ^{2 \times\mathbb{Q}_{p}^{\times}/\left[ \mathbb{Q}_{p}^{\times}\right] ^{2}$ into $\left\{ 1,-1\right\} $. For a fixed $\beta\in\mathbb{Q}_{p}^{\times}$, ${\LARGE \pi}_{\beta}\left( t\right) =\left( \beta,t\right) _{p}$ defines a multiplicative character on $\mathbb{Q}_{p}^{\times}$, the multiplicatively of ${\LARGE \pi}_{\beta}$ follows from property (\ref{HS_2}). \subsection{The Weil Constant} Le \begin{equation} f(x):=a_{1}x_{1}^{2}+a_{2}x_{2}^{2}+\cdots+a_{n}x_{n}^{2},\quad a_{i \in\mathbb{Q}_{p}^{\times},\quad i=1,2,\dots,n, \label{cuadra \end{equation} be a \textit{quadratic form}. A such quadratic form is characterized by three invariants: (i) the dimension $n$; (ii) the discriminant $D=a_{1}a_{2}\cdots a_{n}\operatorname{mod}\left[ \mathbb{Q}_{p}^{\times}\right] ^{2}$; (iii) the Hasse invariant ${H=\prod_{i<j}(a_{i},a_{j})_{p}}$. By \cite[Theoreme 2]{We2}, see also \cite[Theoreme 1.1]{R-S}, there exist a complex constant $\gamma(f)$ of absolute value one, such that \begin{align} & \int_{\mathbb{Q}_{p}^{n}}\hat{\varphi}(x)\chi(tf(x))d^{n}x\nonumber\\ & =\gamma(tf)|t|_{p}^{-n/2}|D|_{p}^{-1/2}\int_{\mathbb{Q}_{p}^{n} \varphi(x)\chi\left( -\frac{1}{t}f{\left( \frac{x_{1}}{2a_{1}},\dots ,\frac{x_{n}}{2a_{n}}\right) }\right) d^{n}x, \label{weil_fun_eq \end{align} for all $t\in\mathbb{Q}_{p}^{\times}$, where $D=a_{1}a_{2}\cdots a_{n}$. Since $\gamma(f)=\gamma(a_{1}x_{1}^{2})\cdots\gamma(a_{n}x_{n}^{2})$, see e.g. \cite[p. 173]{We2}, the calculation of $\gamma(f)$ is reduced to the case $n=1$. For a $\alpha\in\mathbb{Q}_{p}^{\times}$, we set $\gamma(\alpha ):=\gamma(\alpha x_{1}^{2})$. \begin{lemma} \label{lemma0}For a unit $u\in\mathbb{Z}_{p}^{\times}$, with $u\equiv u_{0}\operatorname{mod}p\mathbb{Z}_{p}$, we have $\gamma(u)=1$ and $\gamma(up)=\left( \frac{u_{0}}{p}\right) \sigma_{p}$, where \begin{equation} {\sigma_{p}: \begin{cases} 1 & \text{if }p\equiv1\operatorname{mod}4\\ \sqrt{-1} & \text{if }p\equiv3\operatorname{mod}4. \end{cases} } \label{sigma_p \end{equation} \end{lemma} \begin{proof} Take $\varphi(x)$ to be the characteristic function of $\mathbb{Z}_{p}$ and $u\in\mathbb{Z}_{p}^{\times}$, by (\ref{weil_fun_eq}) \begin{align*} \int_{\mathbb{Q}_{p}}\hat{\varphi}(x)\chi(ux^{2})dx & =\gamma(u)|u|_{p ^{-1/2}\int_{\mathbb{Q}_{p}}\varphi(x)\chi\left( -\frac{x^{2}}{4u}\right) dx\\ \int_{\mathbb{Z}_{p}}dx & =\gamma(u). \end{align*} In the case $up$ with $u\in\mathbb{Z}_{p}^{\times}$, by applying (\ref{weil_fun_eq}) we hav \begin{align*} \int_{\mathbb{Q}_{p}}\hat{\varphi}(x)\chi(upx^{2})dx & =\gamma (up)|up|_{p}^{-1/2}\int_{\mathbb{Q}_{p}}\varphi(x)\chi\left( -\frac{x^{2 }{4pu}\right) dx\\ \int_{\mathbb{Z}_{p}}\chi(upx^{2})dx & =\gamma(up)p^{1/2}\int_{\mathbb{Z _{p}}\chi\left( -\frac{x^{2}}{4pu}\right) dx\\ 1 & =\gamma(up)p^{1/2}\int_{\mathbb{Z}_{p}}\chi\left( -\frac{x^{2} {4pu}\right) dx. \end{align*} If $z\in\mathbb{Z}_{p}\smallsetminus\left\{ 0\right\} $ we set $z=z_{0}+z_{1}p+\ldots+z_{k}p^{k}+\ldots$ with $z_{k}=\left\{ 0,1,\ldots ,p-1\right\} $. Now by changing variables ($x=2uy$) in the previous integral: \begin{align*} \int_{\mathbb{Z}_{p}}\chi\left( -\frac{x^{2}}{4pu}\right) dx & =\int_{|y|_{p}\leq1}\chi\left( -\frac{uy^{2}}{p}\right) dy\\ & =\int_{|y|_{p}=1}\chi\left( -\frac{uy^{2}}{p}\right) dy+\int_{|y|_{p <1}\chi\left( -\frac{uy^{2}}{p}\right) dy\\ & =\frac{1}{p}\sum_{y_{0}=1}^{p-1}exp\left\{ -2\pi i\frac{u_{0}y_{0}^{2} {p}\right\} +\frac{1}{p}=\frac{1}{p}\sum_{y_{0}=0}^{p-1}exp\left\{ -2\pi i\frac{u_{0}y_{0}^{2}}{p}\right\} \\ & =p^{-1/2}\left( \frac{-u_{0}}{p}\right) \sigma_{p}, \end{align*} where in the last step we used a result of Gauss on quadratic exponential sums, see e.g. \cite[p. 55]{V-V-Z}. Therefore \[ \gamma(up)=\frac{1}{\left( \frac{-1}{p}\right) \left( \frac{u_{0} {p}\right) \sigma_{p}}=\left( \frac{u_{0}}{p}\right) \sigma_{p}. \] \end{proof} The next lemma shows the relation between the constant $\gamma$ and the Hilbert symbol. \begin{lemma} \label{lemma1}With the above notation, the following assertions hold. (i) $\gamma(-a)\gamma(a)=1$. (ii) Set $h(x)=x_{1}^{2}-ax_{2}^{2}-bx_{3}^{2}+abx_{4}^{2}$ with $a,b\in\mathbb{Q}_{p}^{\times}$. Then \[ \gamma(h)=\gamma(1)\gamma(-a)\gamma(-b)\gamma(ab)=(a,b)_{p}. \] (iii) If $n\equiv0\operatorname{mod}2$, then $\gamma(tf)=\gamma(f)(t,D^{\ast })_{p}$ for any $t\in\mathbb{Q}_{p}^{\times}$, where \[ D^{\ast}:=(-1)^{\frac{n}{2}}D. \] \end{lemma} \begin{proof} (i) See \cite[Section No. 25, p. 173 ]{We2}. (ii) See \cite[Section No. 28, p. 176]{We2}. (iii) See \cite[Proposition 1.7]{R-S}. \end{proof} \subsection{Local zeta functions} For $a>0$ and $s\in\mathbb{C}$ we set $a^{s}:=e^{s\ln a}$. Let $f(x)$ be a quadratic form over $\mathbb{Q}_{p}$ and ${\LARGE \pi}_{\beta}(t):=(\beta ,t)_{p},\quad t\in\mathbb{Q}_{p}^{\times}$ as before. The local function zeta attached to $\left( f,{\LARGE \pi}_{\beta}\right) $ is the distribution given by \begin{equation} Z_{\varphi}(s,{\LARGE \pi}_{\beta},f):=Z_{\varphi}(s,{\LARGE \pi}_{\beta })=\int_{\mathbb{Q}_{p}^{n}\smallsetminus f^{-1}(0)}{\LARGE \pi}_{\beta }(f(x))|f(x)|_{p}^{s-n/2}\varphi(x)d^{n}x\text{, \ }\label{zeatfunction \end{equation} $\varphi\in\mathbf{S}\left( \mathbb{Q}_{p}^{n}\right) $\ and $\operatorname{Re}(s)>\frac{n}{2}$. If $\beta=1$ we use $Z_{\varphi}(s,f)$ instead of $Z_{\varphi}(s,{\LARGE \pi}_{1},f)$. The local zeta functions are defined for arbitrary polynomials and arbitrary multiplicative characters. These objects were introduced in the 60's by A. Weil and since then they have been studied intensively, see e.g. \cite{Igusa}. The local zeta function $Z_{\varphi}(s,{\LARGE \pi}_{\beta})$ is a distribution on $\mathbf{S}\left( \mathbb{Q}_{p}^{n}\right) $\ for $\operatorname{Re}(s)>\frac{n}{2}$, which admits a meromorphic continuation to the whole complex plane (for arbitrary $f$ and ${\LARGE \pi}_{\beta}$) such that $Z_{\varphi}(s,{\LARGE \pi}_{\beta })$ is a rational function of $p^{-s}$, see \cite[Theorem 8.2.1]{Igusa}. \subsection{Functional equations} It is well-known that the Fourier transform of the distribution ${\LARGE \pi }_{\beta}(t)|t|_{p}^{s-1}$ is $\rho({\LARGE \pi}_{\beta},s){\LARGE \pi _{\beta}^{-1}(t)|t|_{p}^{-s}$ i.e. \begin{equation} \int_{\mathbb{Q}_{p}^{\times}}\widehat{\varphi}(t){\LARGE \pi}_{\beta }(t)|t|_{p}^{s-1}dt=\rho({\LARGE \pi}_{\beta},s)\int_{\mathbb{Q}_{p}^{\times }\varphi(t){\LARGE \pi}_{\beta}^{-1}(t)|t|_{p}^{-s}dt, \label{tate1 \end{equation} for all $\varphi(t)\in S(\mathbb{Q}_{p})$, see e.g. \cite[Section VIII.2]{V-V-Z}. We recall that (\ref{tate1}) is a particular case of the functional equation for the Iwasawa-Tate local zeta function see e.g. \cite[Theoerem 2.4.1 and Lemma 2.4.3]{Tate}. We now compute the factors $\rho(\pi_{\beta},s)$ appearing in (\ref{tate1}). \begin{lemma} \label{funro} Set $\mathbb{Q}_{p}^{\times}/\left[ \mathbb{Q}_{p}^{\times }\right] ^{2}=$ $\left\{ 1,\epsilon,p,\epsilon p\right\} $ where $\epsilon $\ is unit which is not square. Then (i) ${\rho(}{\LARGE \pi}{_{1},}${$s$}${)=\frac{1-p^{s-1}}{1-p^{-s}}};$ (ii) ${\rho(}{\LARGE \pi}{_{\epsilon},}${$s$}${)=\frac{1+p^{s-1}}{1+p^{-s}}}$; (iii) ${\rho(}{\LARGE \pi}{_{\eta},}${$s$}${)=\pm\sigma_{p}p^{s-\frac{1}{2}} $, $\eta=p,\epsilon p$ with ${\sigma_{p}}$\ as in (\ref{sigma_p}). \end{lemma} \begin{proof} (i) Take $\varphi(t)$ to be the characteristic function of $\mathbb{Z}_{p}$ in (\ref{tate1}), the \[ \rho(\pi_{1},s)=\frac{\int_{\mathbb{Z}_{p}\smallsetminus\left\{ 0\right\} }|t|_{p}^{s-1}dt}{\int_{\mathbb{Z}_{p}\smallsetminus\left\{ 0\right\} }|t|_{p}^{-s}dt}=\frac{1-p^{s-1}}{1-p^{-s}}. \] (ii) Note that ${\LARGE \pi}_{\epsilon}(t)=(-1)^{ord(t)}$, see \cite[Lemma on p. 130]{V-V-Z}, by taking $\varphi(t)$ to be the characteristic function of $\mathbb{Z}_{p}$ in (\ref{tate1}), we have \[ \rho({\LARGE \pi}_{\epsilon},s)=\frac{1+p^{s-1}}{1+p^{-s}}. \] (iii) Set \[ \mathbb{Q}_{p,\eta}^{\times}:=\left\{ x\in\mathbb{Q}_{p}^{\times}\mid x=a^{2}-\eta b^{2},\quad a,b\in\mathbb{Q}_{p}\right\} \] and \[ sgn_{\eta}(x): \begin{cases} 1 & \text{if $x\in\mathbb{Q}_{p,\eta}^{\times}$}\\ -1 & \text{if $x\notin\mathbb{Q}_{p,\eta}^{\times}.$ \end{cases} \] In \cite[p. 129]{V-V-Z} is proved that$\ {\rho(}{\LARGE \pi}{_{\eta},}${$s }${)=\pm\sqrt{sgn_{\eta}(-1)}p^{s-\frac{1}{2}}}$ for $\eta=p,\epsilon p$. Since $(t,\eta)_{p}=sgn_{\eta}(t)$ we have \[ {\pm\sqrt{sgn_{\eta}(-1)}=\pm\sqrt{(\eta,-1)_{p}}=\pm\sqrt{\left( \frac {-1}{p}\right) }=\pm\sigma_{p}. \] \end{proof} Set ${f^{\ast}(x):=f\left( \frac{x_{1}}{a_{1}},\dots,\frac{x_{n}}{a_{n }\right) }${, and} \[ Z_{\varphi}^{\ast}(s,{\LARGE \pi}_{\beta}):=\int_{\mathbb{Q}_{p ^{n}\smallsetminus f^{\ast-1}(0)}{\LARGE \pi}_{\beta}(f^{\ast}(x))|f^{\ast }(x)|_{p}^{s-n/2}\varphi(x)d^{n}x. \] \begin{theorem} [{\cite[Theoreme 22-13]{R-S}}]\label{funcional} If $n\equiv0\operatorname{mod 2$, then $Z_{\varphi}(s)$ satisfies \[ Z_{\hat{\varphi}}(s)=\rho({\LARGE \pi}_{1},s-\frac{n}{2}+1)\rho({\LARGE \pi }_{D^{\ast}},s)|D|_{p}^{-1/2}\gamma(f)Z_{\varphi}^{\ast}(-s+n/2,{\LARGE \pi }_{D^{\ast}}) \] for any $\varphi\in\mathbf{S}\left( \mathbb{Q}_{p}^{n}\right) $. \end{theorem} \begin{proof} The announced formula is a particular case of formula 2-20 in \cite{R-S}. We note that our functional equation equals up to a constant to the functional equation in \cite{R-S}, this is due fact that we used a different normalization for the Haar measure. \end{proof} \subsection{Some explicit functional equations} \begin{corollary} \label{Cor_Fun_Eq}Let $f(x)$ be as before. Assume that $n\equiv 0\operatorname{mod}2$ and that $D^{\ast}$ is a square. The \[ Z_{\hat{\varphi}}(s)=\rho({\LARGE \pi}_{1},s-\frac{n}{2}+1)\rho({\LARGE \pi }_{1},s)|D|_{p}^{-1/2}\gamma(f)Z_{\varphi}^{\ast}(-s+n/2) \] for any $\varphi\in\mathbf{S}\left( \mathbb{Q}_{p}^{n}\right) $. \end{corollary} \begin{proposition} \label{eqfun2} If $f(x)=x_{1}^{2}-\eta x_{2}^{2}$, $\eta=\epsilon,p,p\epsilon $, then \begin{align*} & {\displaystyle\int\limits_{\mathbb{Q}_{p}^{2}\smallsetminus\left\{ 0\right\} }} |f(x)|_{p}^{s-1}\widehat{\varphi}(x)d^{2}x\\ & \begin{cases} \dfrac{1-p^{2(s-1)}}{1-p^{-2s} {\displaystyle\int\limits_{\mathbb{Q}_{p}^{2}\smallsetminus\left\{ 0\right\} }} {|\eta x_{1}^{2}-x_{2}^{2}|_{p}^{-s}\varphi(x)d}^{2}x & \text{if \eta=\epsilon\\ \dfrac{1-p^{s-1}}{1-p^{-s} {\displaystyle\int\limits_{\mathbb{Q}_{p}^{2}\smallsetminus\left\{ 0\right\} }} {|\eta x_{1}^{2}-x_{2}^{2}|_{p}^{-s}\varphi(x)d}^{2}x & \text{if \eta=p,p\epsilon. \end{cases} \end{align*} \end{proposition} \begin{proof} Since $D^{\ast}=-D=\eta$ and ${\LARGE \pi}_{-D}(f^{\ast}(x))=(\eta,x_{1 ^{2}-\eta x_{2}^{2})_{p}=1$. By Theorem \ref{funcional}, we have \[ Z_{\widehat{\varphi}}(s)=\rho({\LARGE \pi}_{1},s)\rho({\LARGE \pi _{-D},s)|D|_{p}^{-1/2}\gamma(f)Z_{\varphi}^{\ast}(1-s,{\LARGE \pi}_{1}) \ \[ =\rho({\LARGE \pi}_{1},s)\rho({\LARGE \pi}_{-D},s)|D|_{p}^{-1/2}\gamma (f)|\eta|_{p}^{s {\displaystyle\int\limits_{\mathbb{Q}_{p}^{2}\smallsetminus\left\{ 0\right\} }} {|\eta x_{1}^{2}-x_{2}^{2}|_{p}^{-s}\varphi\left( x\right) dx}_{{1} {dx}_{{2}}. \] The announced functional equations follow from the following calculations. (i) Take $\eta=\epsilon$, then $|\epsilon|_{p}^{-1/2}=|\epsilon|_{p}^{s =1,\quad\gamma(f)=\gamma(1)\gamma(-\epsilon)=1$, see Lemma \ref{lemma0}, and ${\LARGE \pi}_{-D}\left( \epsilon\right) =\left( \epsilon,\epsilon\right) _{p}=1$, see (\ref{HS_3}). Furthermore ${\rho(}{\LARGE \pi}{_{1},}${$s$ ${){=}\frac{1-p^{s-1}}{1-p^{-s}}}$, ${{\rho(}{\LARGE \pi}{_{\epsilon},s)= }\frac{1+p^{s-1}}{1+p^{-s}}$, see Lemma \ref{funro}. (ii) Take $\eta=p$, $p\epsilon$, in this case we have $|\eta|_{p ^{-1/2}=p^{1/2}$, $\gamma(f)=\gamma(-\eta)=\frac{1}{\pm\sigma_{p}}$ (see Lemma \ref{lemma0}) and ${\rho(}{\LARGE \pi}{_{\eta},s)=\pm\sigma_{p}p^{s-\frac {1}{2}}}$ (see Lemma \ref{funro}). Then \begin{align*} \rho({\LARGE \pi}_{1},s)\rho({\LARGE \pi}_{-D},s)|D|_{p}^{-1/2}\gamma (f)|\eta|_{p}^{s} & ={\frac{1-p^{s-1}}{1-p^{-s}}}(\pm\sigma_{p})p^{s-\frac {1}{2}}p^{1/2}(\frac{1}{\pm\sigma_{p}})p^{-s}\\ & =\frac{1-p^{s-1}}{1-p^{-s}}. \end{align*} \end{proof} \begin{proposition} \label{eqfun4} Take $f(x)=x_{1}^{2}-ax_{2}^{2}-px_{3}^{2}+apx_{4}^{2}$, with $a\in\mathbb{Z}$ a quadratic non-residue module $p$. Then \[ \int\limits_{\mathbb{Q}_{p}^{4}\setminus\{0\}}|f(x)|^{s-2}\widehat{\varphi }(x)d^{4}x=\frac{1-p^{s-2}}{\left( 1-p^{-s}\right) }\int\limits_{\mathbb{Q _{p}^{4}\setminus\{0\}}|apx_{1}^{2}-px_{2}^{2}-ax_{3}^{2}+x_{4}^{2}|_{p ^{-s}\varphi(x)d^{4}x. \] \end{proposition} \begin{proof} In this case $n=4$, $D=p^{2}a^{2}$, $D^{\ast}=D$ and $\gamma(f)=(a,p)_{p}=-1$, see Lemma \ref{lemma1} (ii) and (\ref{HS_3}), the functional equation takes the form \begin{align*} & \int\limits_{\mathbb{Q}_{p}^{4}\setminus\{0\}}|f(x)|_{p}^{s-2 \widehat{\varphi}(x)d^{4}x\\ & =-\rho({\LARGE \pi}_{1},s-1)\rho({\LARGE \pi}_{1},s)p\int \limits_{\mathbb{Q}_{p}^{4}\setminus\{0\}}|x_{1}^{2}-a^{-1}x_{2}^{2 -p^{-1}x_{3}^{2}+(ap)^{-1}x_{4}^{2}|_{p}^{-s}\varphi(x)d^{4}x\\ & =-|ap|_{p}^{s}p\frac{1-p^{s-2}}{1-p^{1-s}}\frac{1-p^{s-1}}{1-p^{-s} \int\limits_{\mathbb{Q}_{p}^{4}\setminus\{0\}}|apx_{1}^{2}-px_{2}^{2 -ax_{3}^{2}+x_{4}^{2}|_{p}^{-s}\varphi(x)d^{4}x\\ & =\frac{1-p^{s-2}}{\left( 1-p^{-s}\right) }\int\limits_{\mathbb{Q}_{p ^{4}\setminus\{0\}}|apx_{1}^{2}-px_{2}^{2}-ax_{3}^{2}+x_{4}^{2}|_{p ^{-s}\varphi(x)d^{4}x. \end{align*} \end{proof} \section{\label{Section2}Riesz Kernels and Lizorkin Spaces of Second Kind} In this section we introduce a new type of Riesz kernels depending on a complex parameter and a certain quadratic form. The main result of this section establishes that these kernels considered as distributions on a Lizorkin spaces of second kind form an Abelian group under convolution. \subsection{Riesz Kernels} In this section $f(x):=x_{1}^{2}-ax_{2}^{2}-px_{3}^{2}+apx_{4}^{2}$ with $a\in\mathbb{Z}$ a quadratic non-residue module $p$. Note that $a\in \mathbb{Z}^{\times}_{p}$ and that $f(x)$ is an elliptic quadratic form, i.e. $f(x)=0\Leftrightarrow x=0$. We call the function \[ K_{\alpha}(x):=\dfrac{1-p^{-\alpha}}{1-p^{\alpha-2}}|f(x)|_{p}^{\alpha -2},\quad\operatorname{Re}(\alpha)>0,\ \alpha\neq2+\frac{2\pi\sqrt{-1}}{\ln p}\mathbb{Z}, \] \textit{the Riesz kernel attached to} $f(x)$. \begin{lemma} \label{lemma2}Set $\varphi$ to be the characteristic function of the ball $\widetilde{x}_{0}+\left( p^{m}\mathbb{Z}_{p}\right) ^{4}$. The \[ Z_{\varphi}\left( \alpha,f\right) =\left\{ \begin{array} [c]{lll \frac{p^{-2\alpha m}\left( 1-p^{-2}\right) }{1-p^{-\alpha}} & \text{if} & \widetilde{x}_{0}\in\left( p^{m}\mathbb{Z}_{p}\right) ^{4}\\ & & \\% \begin{array} [c]{c \text{holomorphic function}\\ \text{in }\alpha\text{, for }\alpha\in\mathbb{C \end{array} & \text{if} & \widetilde{x}_{0}\notin\left( p^{m}\mathbb{Z}_{p}\right) ^{4}. \end{array} \right. \] \end{lemma} \begin{proof} We consider first the case $\widetilde{x}_{0}\in\left( p^{m}\mathbb{Z _{p}\right) ^{4}$. Set \[ Z(\alpha): {\displaystyle\int\limits_{\mathbb{Z}_{p}^{4}\smallsetminus\{0\}}} \left\vert f(x)\right\vert _{p}^{\alpha-2}d^{4}x\text{ for }\operatorname{Re (\alpha)>2. \] By a change of variables $Z_{\varphi}\left( \alpha,f\right) =p^{-2\alpha m}Z(\alpha)$. The result follows from the following formula: \begin{equation} Z(\alpha)=\frac{1-p^{-2}}{1-p^{-\alpha}}\text{ for }\operatorname{Re (\alpha)>2. \label{formula \end{equation} Set $\mathbb{Z}_{p}^{4}=\left( p\mathbb{Z}_{p}\right) ^{4 {\textstyle\bigsqcup} U$ with $U=\left\{ x\in\mathbb{Z}_{p}^{4}:\left\Vert x\right\Vert _{p}=1\right\} $. The \begin{align*} Z(\alpha) & {\displaystyle\int\limits_{\left( p\mathbb{Z}_{p}\right) ^{4}}} \left\vert f(x)\right\vert _{p}^{\alpha-2}d^{4}x {\displaystyle\int\limits_{U}} \left\vert f(x)\right\vert _{p}^{\alpha-2}d^{4}x\\ & =p^{-2\alpha}Z(\alpha) {\displaystyle\int\limits_{U}} \left\vert f(x)\right\vert _{p}^{\alpha-2}d^{4}x, \end{align*} i.e. \[ Z(\alpha)=\frac{1}{1-p^{-2\alpha} {\displaystyle\int\limits_{U}} \left\vert f(x)\right\vert _{p}^{\alpha-2}d^{4}x\text{. \] In order to show (\ref{formula}), it is sufficient to prove the following formula: \begin{equation {\displaystyle\int\limits_{U}} \left\vert f(x)\right\vert _{p}^{\alpha-2}d^{4}x=(1-p^{-2})\left( 1+p^{-\alpha}\right) \text{ for }\alpha\in\mathbb{C}. \label{Int_unidades \end{equation} This formula can be established as follows. For $i=\left( i_{1},i_{2 ,i_{3},i_{4}\right) \in\left\{ 0,1\right\} ^{4}\smallsetminus\left\{ \left( 1,1,1,1\right) \right\} $ we define \begin{align*} U^{\left( i\right) } & =U_{1}^{\left( i\right) }\times U_{2}^{\left( i\right) }\times U_{3}^{\left( i\right) }\times U_{4}^{\left( i\right) },\\ U_{j}^{\left( i\right) } & :=\left\{ \begin{array} [c]{ccc p\mathbb{Z}_{p} & \text{if} & i_{j}=1\\ & & \\ \mathbb{Z}_{p}^{\times} & \text{if} & i_{j}=0. \end{array} \right. \end{align*} Then $U {\textstyle\bigsqcup\nolimits_{i}} U^{\left( i\right) }$ and \ {\displaystyle\int\limits_{U}} \left\vert f(x)\right\vert _{p}^{\alpha-2}d^{4}x {\displaystyle\sum\limits_{i}} {\displaystyle\int\limits_{U^{\left( i\right) }}} \left\vert f(x)\right\vert _{p}^{\alpha-2}d^{4}x: {\displaystyle\sum\limits_{i}} Z_{i}(\alpha). \] By a direct calculation one finds \ \begin{tabular} [c]{|l|l|}\hline Index $i$ & $Z_{i}(\alpha)$\\\hline \begin{array} [c]{c \left( 1,1,1,0\right) \end{array} $ & $p^{-\alpha-1}\left( 1-p^{-1}\right) $\\\hline \begin{array} [c]{c \left( 1,1,0,1\right) \end{array} $ & $p^{-\alpha-1}\left( 1-p^{-1}\right) $\\\hline \begin{array} [c]{c \left( 1,1,0,0\right) \end{array} $ & $p^{-\alpha}\left( 1-p^{-1}\right) ^{2}$\\\hline \begin{array} [c]{cc \left( 1,0,1,1\right) , & \left( 0,1,1,1\right) \end{array} $ & $\left( 1-p^{-1}\right) p^{-3}$\\\hline \begin{array} [c]{ccc \left( 1,0,1,0\right) , & \left( 1,0,0,1\right) , & \left( 0,1,1,0\right) \\ \left( 0,1,0,1\right) , & \left( 0,0,1,1\right) & \end{array} $ & $\left( 1-p^{-1}\right) ^{2}p^{-2}$\\\hline \begin{array} [c]{ccc \left( 1,0,0,0\right) , & \left( 0,1,0,0\right) , & \left( 0,0,1,0\right) \\ \left( 0,0,0,1\right) & & \end{array} $ & $\left( 1-p^{-1}\right) ^{3}p^{-1}$\\\hline \begin{array} [c]{c \left( 0,0,0,0\right) \end{array} $ & $\left( 1-p^{-1}\right) ^{4}$\\\hline \end{tabular} \ \ \ \ \ \] In the case $\widetilde{x}_{0}\notin\left( p^{m}\mathbb{Z}_{p}\right) ^{4}$, $f$\ does not vanish on the ball $\widetilde{x}_{0}+\left( p^{m \mathbb{Z}_{p}\right) ^{4}$ which implies that $Z_{\varphi}\left( \alpha\right) $ is a holomorphic function on the whole complex plane. \end{proof} \begin{lemma} \label{lemma3}$K_{\alpha}(x)$ possesses, as a distribution on $S(\mathbb{Q _{p}^{4})$, a meromorphic continuation to all $\alpha\neq2+\frac{2\pi\sqrt {-1}}{\ln p}\mathbb{Z}$ given by \begin{multline*} \left\langle K_{\alpha},\varphi\right\rangle =\varphi(0)\dfrac{1-p^{-2 }{1-p^{\alpha-2}}+\\ \dfrac{1-p^{-\alpha}}{1-p^{\alpha-2}}\left[ \int_{||x||_{p}>1}\varphi (x)|f(x)|_{p}^{\alpha-2}d^{4}x+\int_{||x||_{p}\leq1}\left( \varphi (x)-\varphi\left( 0\right) \right) |f(x)|_{p}^{\alpha-2}d^{4}x\right] . \end{multline*} \end{lemma} \begin{proof} The result follows from Lemma \ref{lemma2}\ \ b \begin{multline*} \left\langle K_{\alpha},\varphi\right\rangle =\dfrac{1-p^{-\alpha }{1-p^{\alpha-2}}\int\limits_{\mathbb{Q}_{p}^{4}\setminus\{0\}}\varphi (x)|f\left( x\right) |_{p}^{\alpha-2}d^{4}x=\\ \dfrac{1-p^{-\alpha}}{1-p^{\alpha-2}}\left[ \int_{||x||_{p}>1}\varphi (x)|f\left( x\right) |_{p}^{\alpha-2}d^{4}x+\int_{||x||_{p}\leq1}\left( \varphi(x)-\varphi\left( 0\right) \right) |f(x)|_{p}^{\alpha-2 d^{4}x\right] \\ +\varphi(0)\dfrac{1-p^{-2}}{1-p^{\alpha-2}}. \end{multline*} \end{proof} From Lemma \ref{lemma3} follows that the distribution $K_{\alpha}$ has simple poles at the points $\alpha=2+\alpha_{k}$ with $\alpha_{k}:=\frac{2k\pi \sqrt{-1}}{\ln p},\quad k\in\mathbb{Z}$, and \begin{equation} {\lim_{\alpha\rightarrow\alpha_{k}}K_{\alpha}:=K_{\alpha_{k}}=\delta,} \label{deltacero \end{equation} where ${\delta}$ denotes the Dirac distribution. \begin{lemma} \label{lemma4 \[ \int_{||x||_{p}>1}\frac{1}{|f\left( x\right) |_{p}^{\alpha+2} dx=\frac{p^{-2\alpha}(1-p^{-2})\left( 1+p^{\alpha}\right) }{1-p^{-2\alpha },\quad\text{Re}(\alpha)>0. \] \end{lemma} \begin{proof} Set $U=(\mathbb{Z}_{p})^{4}\setminus(p\mathbb{Z}_{p})^{4}$ as before. Then \begin{multline*} \int_{||x||_{p}>1}\frac{1}{|f\left( x\right) |_{p}^{\alpha+2}}d^{4 x=\sum_{m=1}^{\infty}\int_{p^{-m}U}\frac{1}{|f\left( x\right) |^{\alpha+2 }d^{4}x\\ =\sum_{m=1}^{\infty}p^{-2m\alpha}\int_{U}\frac{1}{|f\left( x\right) |_{p}^{\alpha+2}}d^{4}x=\frac{p^{-2\alpha}}{1-p^{-2\alpha}}\int_{U}\frac {1}{|f\left( x\right) |_{p}^{\alpha+2}}d^{4}x\\ =\frac{p^{-2\alpha}(1-p^{-2})\left( 1+p^{\alpha}\right) }{1-p^{-2\alpha}}, \end{multline*} where we used (\ref{Int_unidades}). \end{proof} \begin{proposition} \label{Prop1}For $\text{Re}(\alpha)>0$ and $\varphi\in\mathbf{S}\left( \mathbb{Q}_{p}^{4}\right) $, the following formulas hold: (i) $\left\langle K_{\alpha},\varphi\right\rangle =\dfrac{1-p^{-\alpha }{1-p^{\alpha-2} {\displaystyle\int\nolimits_{\mathbb{Q}_{p}^{4}\setminus\{0\}}} |f\left( x\right) |_{p}^{\alpha-2}\varphi(x)d^{4}x,\quad\alpha\neq 2+\alpha_{k}$; (ii) $\left\langle K_{-\alpha},\varphi\right\rangle =\dfrac{1-p^{\alpha }{1-p^{-\alpha-2} {\displaystyle\int\nolimits_{\mathbb{Q}_{p}^{4}}} \frac{\varphi(x)-\varphi(0)}{|f\left( x\right) |_{p}^{\alpha+2}}d^{4}x$; (iii) $(K_{\alpha}\ast\varphi)(x)=\dfrac{1-p^{-\alpha}}{1-p^{\alpha-2} {\displaystyle\int\nolimits_{\mathbb{Q}_{p}^{4}\setminus\{0\}}} |f(y)|_{p}^{\alpha-2}\varphi(x+y)d^{4}y,\quad\alpha\neq2+\alpha_{k}$; (iv) $(K_{-\alpha}\ast\varphi)(x)=\dfrac{1-p^{\alpha}}{1-p^{-\alpha-2} {\displaystyle\int\nolimits_{\mathbb{Q}_{p}^{4}}} \frac{\varphi(x+y)-\varphi(x)}{|f(y)|_{p}^{\alpha+2}}d^{4}y.$ \end{proposition} \begin{proof} (i) Since every test function can be written a finite sums of characteristic functions of balls, Lemma \ref{lemma2} implies that \[ \dfrac{1-p^{-\alpha}}{1-p^{\alpha-2} {\displaystyle\int\limits_{\mathbb{Q}_{p}^{4}\setminus\{0\}}} |f\left( x\right) |_{p}^{\alpha-2}\varphi(x)d^{4}x \] is well-defined for $\operatorname{Re}(\alpha)>0$ and $\alpha\neq2+\alpha_{k $. The announced formula follows by a calculation similar to the one done in the proof of Lemma \ref{lemma3}. (ii) We first note that the integra \[ \dfrac{1-p^{\alpha}}{1-p^{-\alpha-2} {\displaystyle\int\nolimits_{\mathbb{Q}_{p}^{4}}} \frac{\varphi(x)-\varphi(0)}{|f\left( x\right) |_{p}^{\alpha+2}}d^{4}x \] converges on $\operatorname{Re}(\alpha)>0$. Indeed, since $f\left( x\right) $ is an elliptic quadratic form we hav \begin{equation} B\left\Vert x\right\Vert _{p}^{2}\leq|f\left( x\right) |_{p}\leq A\left\Vert x\right\Vert _{p}^{2}\text{ for any }x\in\mathbb{Q}_{p}^{n ,\label{Zuniga_ineq \end{equation} where $A$, $B$ are positive constants, cf. \cite[Lemma 1]{Z-G3}, then \ {\displaystyle\int\nolimits_{\mathbb{Q}_{p}^{4}}} \frac{\left\vert \varphi(x)-\varphi(0)\right\vert }{|f\left( x\right) |_{p}^{\operatorname{Re}\left( \alpha\right) +2}}d^{4}x\leq\frac{2\left\Vert \varphi\right\Vert _{L^{\infty}}}{B^{\operatorname{Re}(\alpha)+2} {\displaystyle\int\nolimits_{\left\Vert x\right\Vert _{p}>p^{m}}} \frac{1}{\left\Vert x\right\Vert _{p}^{2\operatorname{Re}(\alpha)+4} d^{4}x<\infty, \] where $m$ is the exponent of local constancy of $\varphi$. No \begin{multline*} \dfrac{1-p^{\alpha}}{1-p^{-\alpha-2} {\displaystyle\int\nolimits_{\mathbb{Q}_{p}^{4}}} \frac{\varphi(x)-\varphi(0)}{|f\left( x\right) |_{p}^{\alpha+2}}d^{4}x=\\ \dfrac{1-p^{\alpha}}{1-p^{-\alpha-2}}\left\{ \int_{||x||_{p}\leq1 \frac{\varphi(x)-\varphi(0)}{|f\left( x\right) |_{p}^{2+\alpha}}d^{4 x+\int_{||x||_{p}>1}\frac{\varphi(x)}{|f\left( x\right) |_{p}^{2+\alpha }d^{4}x\right\} -\varphi(0)\dfrac{1-p^{\alpha}}{1-p^{-\alpha-2}}\\ \times\int_{||x||_{p}>1}\frac{1}{|f\left( x\right) |_{p}^{2+\alpha}}d^{4}x\\ =\dfrac{1-p^{\alpha}}{1-p^{-\alpha-2}}\left\{ \int_{||x||_{p}\leq1 \frac{\varphi(x)-\varphi(0)}{|f\left( x\right) |_{p}^{2+\alpha}}d^{4 x+\int_{||x||_{p}>1}\frac{\varphi(x)}{|f\left( x\right) |_{p}^{2+\alpha }d^{4}x\right\} \\ -\varphi(0)\dfrac{1-p^{\alpha}}{1-p^{-\alpha-2}}\frac{p^{-2\alpha (1-p^{-2})\left( 1+p^{\alpha}\right) }{1-p^{-2\alpha}}\\ =\left\langle K_{-\alpha},\varphi\right\rangle , \end{multline*} where we used Lemmas \ref{lemma4} and \ref{lemma3}. (iii)-(iv) We recall that if $\varphi\in\mathbf{S}\left( \mathbb{Q}_{p ^{4}\right) $, then $(K_{\alpha}\ast\varphi)(x)=\left\langle K_{\alpha }(y),\varphi(x-y)\right\rangle $, and since $K_{\alpha}(-y)=K_{\alpha}(y)$, we have $(K_{\alpha}\ast\varphi)(x)=\left\langle K_{\alpha}(y),\varphi (x+y)\right\rangle $. Therefore (iii) follows from (i) and (iv) follows from (ii). \end{proof} \subsection{Lizorkin spaces of second kind} Consider the spaces \[ \mathbf{\Psi}:=\mathbf{\Psi}(\mathbb{Q}_{p}^{n})=\{\psi\in\boldsymbol{S (\mathbb{Q}_{p}^{n})\mid\psi(0)=0\} \] and \[ \mathbf{\Phi}:=\mathbf{\Phi}(\mathbb{Q}_{p}^{n})=\{\phi\mid\phi=\mathcal{F [\psi],\ \ \mathbb{\psi}\in\mathbf{\Psi}(\mathbb{Q}_{p}^{n})\}. \] The space $\mathbf{\Phi}$\ is called \textit{the }$p$-adic \textit{Lizorkin space of test functions of second kind}. We equip $\mathbf{\Psi}$ and $\mathbf{\Phi}$ with the topology inherited from $\boldsymbol{S (\mathbb{Q}_{p}^{n})$. Note that $\mathcal{F}:\mathbf{\Psi\rightarrow\Phi}$ is an isomorphism of linear spaces and $\mathcal{F}\left( \mathcal{F}\left[ \mathbf{\Psi}\right] \right) =\mathbf{\Psi}$. Let $\mathbf{\Phi^{\prime}}=\mathbf{\Phi^{\prime}}(\mathbb{Q}_{p}^{n})$ denote the topological dual of the space $\mathbf{\Phi}(\mathbb{Q}_{p}^{n})$. This is space of\textit{ the }$p$\textit{-adic Lizorkin space of distributions of the second kind}. We define the Fourier transform of distributions $J\in\mathbf{\Phi}^{\prime }(\mathbb{Q}_{p}^{n})$ and $G\in\mathbf{\Psi}^{\prime}(\mathbb{Q}_{p}^{n})$ by \begin{align*} \left\langle \mathcal{F}\left[ J\right] ,\psi\right\rangle & =\left\langle J,\mathcal{F}\left[ \psi\right] \right\rangle ,\qquad\text{for any }\psi \in\mathbf{\Psi}(\mathbb{Q}_{p}^{n}),\\ \left\langle \mathcal{F}\left[ G\right] ,\phi\right\rangle & =\left\langle G,\mathcal{F}\left[ \phi\right] \right\rangle ,\qquad\text{\ for any \phi\in\mathbf{\Phi}(\mathbb{Q}_{p}^{n}). \end{align*} It is clear that a $\mathcal{F}[\mathbf{\Psi}^{\prime}(\mathbb{Q}_{p ^{n})]=\mathbf{\Phi}^{\prime}(\mathbb{Q}_{p}^{n})$ and $\mathcal{F [\mathbf{\Phi}^{\prime}(\mathbb{Q}_{p}^{n})]=\mathbf{\Psi}^{\prime (\mathbb{Q}_{p}^{n})$. For further details about $p$-adic Lizorkin spaces the reader may consult \cite{A-K-S}. \subsection{The Riesz kernels form an Abelian group} The goal of this section is to prove the following result: \begin{theorem} \label{mainA}For $\alpha,\beta\in\mathbb{C}$, $K_{\alpha}\ast K_{\beta }=K_{\alpha+\beta}$ in $\mathbf{\Phi}^{\prime}(\mathbb{Q}_{p}^{4})$. \end{theorem} Before giving the proof we need to establish several auxiliary results. \begin{definition} Set $\ f^{\mathbf{\circ}}(x):=apx_{1}^{2}-px_{2}^{2}-ax_{3}^{2}+x_{4}^{2}$. The Riesz kernel attached to $f^{\mathbf{\circ}}(x)$ is the distributio \[ K_{-\alpha}^{\mathbf{\circ}}(x):=|f^{\mathbf{\circ}}(x)|_{p}^{-\alpha}\text{ in }\mathbf{\Psi}^{\prime}(\mathbb{Q}_{p}^{4})\text{, for }\alpha\in \mathbb{C}\text{. \] \end{definition} \begin{proposition} \label{fourietran} Considering $K_{\alpha}\in\mathbf{\Phi}^{\prime (\mathbb{Q}_{p}^{4})$ and $K_{-\alpha}^{\mathbf{\circ}}\in\mathbf{\Psi }^{\prime}(\mathbb{Q}_{p}^{4})$, we hav \[ \mathcal{F}\left[ K_{\alpha}\right] =K_{-\alpha}^{\mathbf{\circ}}\text{ for }\alpha\not =2+\alpha_{k}\text{ and }\alpha\not =\alpha_{k},\text{ k\in\mathbb{Z}\text{. \] \end{proposition} \begin{proof} The formula follows from Proposition \ref{eqfun4}. \end{proof} \begin{lemma} \label{lemma5 \begin{equation} \lim_{\alpha\rightarrow2+\alpha_{k}}\left\langle K_{\alpha},\varphi \right\rangle =-\dfrac{1-p^{-2}}{\ln p}\left\langle \ln|f(x)|_{p ,\varphi(x)\right\rangle \text{ for }\varphi\in\mathbf{\Phi}(\mathbb{Q _{p}^{4}). \label{K_alpha \end{equation} \end{lemma} \begin{remark} We understand the right-hand side in (\ref{K_alpha}) as the distribution induced by the locally integrable function $\ln|f(x)|_{p}:\mathbb{Q}_{p ^{4}\setminus\{0\}\rightarrow\mathbb{R}$. \end{remark} \begin{proof} Sinc \begin{multline*} \lim_{\alpha\rightarrow2+\alpha_{k}}\left\langle K_{\alpha},\varphi \right\rangle =\lim_{\alpha\rightarrow2+\alpha_{k}}\frac{1-p^{-\alpha }{1-p^{\alpha-2} {\displaystyle\int\limits_{\mathbb{Q}_{p}^{4}\smallsetminus\{0\}}} |f(x)|_{p}^{\alpha-2}\varphi(x)d^{4}x\text{ }\\ =\lim_{\beta\rightarrow2}\left( 1-p^{-\beta}\right) {\displaystyle\int\limits_{\mathbb{Q}_{p}^{4}\smallsetminus\{0\}}} \left[ \frac{|x_{1}^{2}-ax_{2}^{2}-px_{3}^{2}+apx_{4}^{2}|_{p}^{\beta-2 -1}{1-p^{\beta-2}}\right] \varphi(x)d^{4}x \end{multline*} by taking $\beta=\alpha-\alpha_{k}$ and by using the fact that $\int \varphi(x)d^{4}x=0$. Now by passing to the limit under the integral sign we hav \[ \lim_{\alpha\rightarrow2+\alpha_{k}}\left\langle K_{\alpha},\varphi \right\rangle =-(1-p^{-2} {\displaystyle\int\limits_{\mathbb{Q}_{p}^{4}\smallsetminus\{0\}}} \frac{\ln|f(x)|_{p}}{\ln p}\varphi(x)d^{4}x. \] The passage to the limit under the integral sign is justified by the Lebesgue Dominated Convergence Theorem and the inequality \[ \left\vert \frac{e^{(\beta-2)\ln|f\left( x\right) |_{p}}-1}{1-e^{(\beta -2)\ln p}}\right\vert \leq C\left\vert \frac{\ln|f(x)|_{p}}{\ln p}\right\vert \text{ for }x\in\text{supp }\varphi\subset\mathbb{Q}_{p}^{4}\setminus \{0\}\text{ and }|\beta-2|\leq1\text{, \] where $C=C(p,\text{supp }\varphi)$ is a positive constant. \end{proof} \begin{definition} We define \[ K_{2+\alpha_{k}}(x)=-\dfrac{1-p^{-2}}{\ln p}\ln|f\left( x\right) |_{p \in\mathbf{\Phi^{\prime}}(\mathbb{Q}_{p}^{4}). \] \end{definition} \begin{lemma} \label{Lemma_Fourier_K2 \[ \left\langle \mathcal{F}\left[ K_{2+\alpha_{k}}\right] ,\varphi\right\rangle =\left\langle K_{-2}^{\mathbf{\circ}},\varphi\right\rangle ,\quad\text{for }\varphi\in\mathbf{\Psi}(\mathbb{Q}_{p}^{4}). \] \end{lemma} \begin{proof} By using the fact that $K_{2+\alpha_{k}}=K_{2}$ and by Proposition \ref{fourietran} we get \[ \left\langle \mathcal{F}\left[ K_{2+\alpha_{k}}\right] ,\varphi\right\rangle =\lim_{\alpha\rightarrow2+\alpha_{k}}\left\langle K_{\alpha},\mathcal{F \left[ \varphi\right] \right\rangle =\lim_{\alpha\rightarrow2}\left\langle K_{\alpha},\mathcal{F}\left[ \varphi\right] \right\rangle =\lim _{\alpha\rightarrow2}\left\langle K_{-\alpha}^{\mathbf{\circ}},\varphi \right\rangle . \] By using $\varphi\left( 0\right) =0$, we have \[ \lim_{\alpha\rightarrow2}\left\langle K_{\alpha}^{\mathbf{\circ} ,\varphi\right\rangle =\lim_{\alpha\rightarrow2 {\displaystyle\int\limits_{\left\Vert x\right\Vert _{p}>p^{m}}} \left\vert f^{\mathbf{\circ}}\left( x\right) \right\vert _{p}^{-\alpha }\varphi\left( x\right) d^{4}x, \] where $m\in\mathbb{Z}$ is the exponent of local constancy of $\varphi$. We now use \ the fact that $f^{\mathbf{\circ}}\left( x\right) $ is an elliptic quadratic form to get \begin{equation} B\left\Vert x\right\Vert _{p}^{2}\leq|f^{\mathbf{\circ}}\left( x\right) |_{p}\leq A\left\Vert x\right\Vert _{p}^{2}\text{ for any }x\in\mathbb{Q _{p}^{n}, \label{Zuniga_ineq2 \end{equation} where $A$, $B$ are positive constants, cf. \cite[Lemma 1]{Z-G3}. Without loss of generality we may assume that $B\leq1$ and that $m>0$. The \[ \left\vert f^{\mathbf{\circ}}\left( x\right) \right\vert _{p ^{-\operatorname{Re}(\alpha)}\left\vert \varphi\left( x\right) \right\vert \leq\frac{\left\vert \varphi\left( x\right) \right\vert {B^{\operatorname{Re}(\alpha)}\left\Vert x\right\Vert _{p}^{2\operatorname{Re (\alpha)}}\leq\frac{\left\vert \varphi\left( x\right) \right\vert }{B^{2+\epsilon}p^{2m\left( 2-\epsilon\right) } \] which is an integrable function on $\left( \mathbb{Q}_{p}^{4}\smallsetminus B_{m}(0)\right) \cap$supp$\varphi$ and $\alpha\in\left( 2-\epsilon ,2+\epsilon\right) $, where $\epsilon$ is a small fixed positive constant. Therefore, by applying the Lebesgue Dominated Convergence Theorem, \[ \lim_{\alpha\rightarrow2}\left\langle K_{\alpha}^{\mathbf{\circ} ,\varphi\right\rangle {\displaystyle\int\limits_{\mathbb{Q}_{p}^{4}}} \left\vert f^{\mathbf{\circ}}\left( x\right) \right\vert _{p}^{-2 \varphi\left( x\right) d^{4}x. \] \end{proof} \begin{proposition} \label{fourietransformtheorem} Considering $K_{\alpha}\in\mathbf{\Phi ^{\prime}(\mathbb{Q}_{p}^{4})$ and $K_{-\alpha}^{\mathbf{\circ} \in\mathbf{\Psi}^{\prime}(\mathbb{Q}_{p}^{4})$ \[ \mathcal{F}\left[ K_{\alpha}\right] =K_{-\alpha}^{\mathbf{\circ}}\text{ for }\alpha\in\mathbb{C}\text{. \] \end{proposition} \begin{proof} The result follows from Proposition \ref{fourietran} and Lemma \ref{Lemma_Fourier_K2}. \end{proof} \begin{lemma} \label{lemma6} For any $\alpha$, $\beta\in\mathbb{C}$, $K_{-\alpha }^{\mathbf{\circ}}\cdot K_{-\beta}^{\mathbf{\circ}}=K_{-(\alpha+\beta )}^{\mathbf{\circ}}$ in $\in\mathbf{\Psi}^{\prime}(\mathbb{Q}_{p}^{4})$. \end{lemma} \begin{proof} It follows from Lemma \ref{producdistributions}. Indeed, the functions $K_{-\alpha}^{\mathbf{\circ}}$ and $K_{-\beta}^{\mathbf{\circ}}$ belong to $L_{loc}^{1}$, and \[ \lim_{\xi\rightarrow0 {\displaystyle\int\limits_{\mathbb{Q}_{p}^{4}}} K_{-\alpha}^{\mathbf{\circ}}(x)\varphi(x)K_{-\beta}^{\mathbf{\circ} (x-\xi)d^{4}x {\displaystyle\int\limits_{\mathbb{Q}_{p}^{4}}} K_{-\alpha}^{\mathbf{\circ}}(x)\varphi(x)K_{-\beta}^{\mathbf{\circ}}(x)d^{4}x. \] This last statement follows from the Lebesgue Dominated Convergence Theorem and (\ref{Zuniga_ineq2}) by the inequality \[ \left\vert K_{-\alpha}^{\mathbf{\circ}}(x)\varphi(x)K_{-\beta}^{\mathbf{\circ }}(x-\xi)\right\vert \leq C\left( \varphi,\alpha,\beta\right) ||x||_{p ^{-2\operatorname{Re}\left( \alpha\right) -2\operatorname{Re}\left( \beta\right) \] for $x\in$supp $\varphi$ and $||\xi||_{p}\leq p^{m\left( \varphi\right) }$, where $C\left( \varphi,\alpha,\beta\right) $ is a positive constant and $m\left( \varphi\right) $ is the largest integer satisfying $\varphi \mid_{B_{m\left( \varphi\right) }(0)}$ $\equiv0$. \end{proof} \subsubsection{Proof of Theorem \ref{mainA}.} By Lemma \ref{lemma6}, for any $\alpha$, $\beta\in\mathbb{C}$, we have \[ \mathcal{F}\left[ K_{-\alpha}^{\mathbf{\circ}}\cdot K_{-\beta}^{\mathbf{\circ }}\right] =\mathcal{F}\left[ K_{-\alpha}^{\mathbf{\circ}}\right] \ast\mathcal{F}\left[ K_{-\beta}^{\mathbf{\circ}}\right] =\mathcal{F}\left[ K_{-\left( \alpha+\beta\right) }^{\mathbf{\circ}}\right] \text{ in }\mathbf{\Phi}^{\prime}(\mathbb{Q}_{p}^{4}). \] By Proposition \ref{fourietransformtheorem}, $K_{\alpha}=\mathcal{F}\left[ K_{-\alpha}^{\mathbf{\circ}}\right] $ for $\alpha\in\mathbb{C}$ since $\mathcal{F}\left[ \mathcal{F}\left[ K_{\alpha}\right] \right] =K_{\alpha }$, therefore for any $\alpha$, $\beta\in\mathbb{C}$, $K_{\alpha}\ast K_{\beta}=K_{\alpha+\beta}$. \begin{remark} \label{nota_dim_2}The proof given Theorem \ref{mainA} can be extended to cover the elliptic quadratic forms of dimension $2$, see Proposition \ref{eqfun2}. \end{remark} \section{\label{Section3}Pseudodifferential Operators and Fundamental Solutions} We take $f(\xi)=\xi_{1}^{2}-a\xi_{2}^{2}-p\xi_{3}^{2}+ap\xi_{4}^{2}$, $f^{\ast}(\xi)=\frac{ap\xi_{1}^{2}-p\xi_{2}^{2}-a\xi_{3}^{2}+\xi_{4}^{2}}{ap $, with $a\in\mathbb{Z}$ a quadratic non-residue module $p$, as in Section \ref{Section2}. Given $\alpha>0$, we define the pseudodifferential operator with symbol $\left\vert f^{\circ}\left( \xi\right) \right\vert _{p}^{\alpha }$ b \ \begin{array} [c]{lll \mathbf{S}\left( \mathbb{Q}_{p}^{4}\right) & \rightarrow & C\left( \mathbb{Q}_{p}^{4}\right) \cap L^{2}\left( \mathbb{Q}_{p}^{4}\right) \\ & & \\ \varphi & \rightarrow & \left( \boldsymbol{f}\left( \partial,\alpha\right) \varphi\right) \left( x\right) :=\mathcal{F}_{\xi\rightarrow x}^{-1}\left( \left\vert f^{\circ}\left( \xi\right) \right\vert _{p}^{\alpha \mathcal{F}_{x\rightarrow\xi}\varphi\right) . \end{array} \] This operator is well-defined since $\left\vert f^{\circ}\left( \xi\right) \right\vert _{p}^{\alpha}\mathcal{F}_{x\rightarrow\xi}\varphi\in L^{1}\left( \mathbb{Q}_{p}^{4}\right) \cap L^{2}\left( \mathbb{Q}_{p}^{4}\right) $. Note that $\left\vert pf^{\ast}\left( \xi\right) \right\vert _{p}^{\alpha }=\left\vert f^{\circ}\left( \xi\right) \right\vert _{p}^{\alpha}$. Since $\mathcal{F}^{-1}\left( \left\vert f^{\circ}\right\vert _{p}^{\alpha}\right) =K_{\alpha}$ (cf. Proposition \ref{fourietransformtheorem} ), by applying Proposition \ref{Prop1} (iv), we get \begin{equation} \boldsymbol{f}\left( \partial,\alpha\right) \varphi=K_{-\alpha}\ast \varphi=\dfrac{1-p^{\alpha}}{1-p^{-\alpha-2} {\displaystyle\int\nolimits_{\mathbb{Q}_{p}^{4}}} \frac{\varphi(x-y)-\varphi(x)}{|f(y)|_{p}^{\alpha+2}}d^{4}y, \label{Ext_oper \end{equation} for $\varphi\in\mathbf{S}\left( \mathbb{Q}_{p}^{4}\right) $. Set $\mathcal{E}_{f,\alpha}\left( \mathbb{Q}_{p}^{4}\right) $ to be the class consisting of locally constant functions $\varphi\left( x\right) $ satisfyin \ {\displaystyle\int\limits_{\left\Vert x\right\Vert _{p}\geq p^{m}}} \frac{\left\vert \varphi(x)\right\vert }{|f(x)|_{p}^{\alpha+2}}d^{4 x<\infty\text{ for some }m\in\mathbb{Z}\text{. \] The operator $\boldsymbol{f}\left( \partial,\alpha\right) $\ can be extended to $\mathcal{E}_{f,\alpha}\left( \mathbb{Q}_{p}^{4}\right) $. \begin{lemma} \label{lemma6A}If $\varphi\in\mathcal{E}_{f,\alpha}\left( \mathbb{Q}_{p ^{4}\right) $, then the integral on the right- hand side of (\ref{Ext_oper}) converges. \end{lemma} \begin{proof} Since $\varphi$ is locally constant there exists $l=l(x)\in\mathbb{Z}$ such that $\varphi(x-y)-\varphi(x)=0$ for $\left\Vert y\right\Vert _{p}\leq p^{l}$, thus, it is sufficient to show the convergence of the following integrals: \ {\displaystyle\int\limits_{\left\Vert y\right\Vert _{p}>p^{l}}} \frac{1}{|f(y)|_{p}^{\alpha+2}}d^{4}y {\displaystyle\int\limits_{\left\Vert y\right\Vert _{p}>p^{l}}} \frac{\left\vert \varphi(x-y)\right\vert }{|f(y)|_{p}^{\alpha+2}}d^{4}y {\displaystyle\int\limits_{\left\Vert x-z\right\Vert _{p}>p^{l}}} \frac{\left\vert \varphi(z)\right\vert }{|f(x-z)|_{p}^{\alpha+2}}d^{4}z. \] The convergence of the first integral follows from (\ref{Zuniga_ineq}). To establish the convergence of the second integral, it is sufficient to \ show the convergence of the integral \begin{equation {\displaystyle\int\limits_{\left\Vert x-z\right\Vert _{p}>p^{l}}} \frac{\left\vert \varphi(z)\right\vert }{\left\Vert x-z\right\Vert _{p}^{2\alpha+4}}d^{4}z, \label{int \end{equation} cf. (\ref{Zuniga_ineq}). The convergence of this last integral is established by considering the cases: (i) $\left\Vert x\right\Vert _{p}<\left\Vert z\right\Vert _{p}$, (ii) $\left\Vert x\right\Vert _{p}>\left\Vert z\right\Vert _{p}$, (iii) $\left\Vert x\right\Vert _{p}=\left\Vert z\right\Vert _{p}$. The verification of cases (i)-(ii) is left to the reader. In the case (iii), we change variables as $x=p^{M}\widetilde{x}$, $z=p^{M}\widetilde{z}$ with $\left\Vert \widetilde{x}\right\Vert _{p}=\left\Vert \widetilde{z}\right\Vert _{p}=1$ in (\ref{int}), the \begin{align* {\displaystyle\int\limits_{\left\Vert x-z\right\Vert _{p}>p^{l}}} \frac{\left\vert \varphi(z)\right\vert }{\left\Vert x-z\right\Vert _{p}^{2\alpha+4}}d^{4}z & =p^{2M\alpha {\displaystyle\int\limits_{\substack{\left\Vert \widetilde{x}-\widetilde {z}\right\Vert _{p}>p^{l+M}\\\left\Vert \widetilde{z}\right\Vert =1}}} \frac{\left\vert \varphi(p^{M}\widetilde{z})\right\vert }{\left\Vert \widetilde{x}-\widetilde{z}\right\Vert _{p}^{2\alpha+4}}d^{4}\widetilde{z}\\ & \leq p^{2M\alpha-\left( 2\alpha+4\right) \left( l+M\right) {\displaystyle\int\limits_{\left\Vert \widetilde{z}\right\Vert =1}} \left\vert \varphi(p^{M}\widetilde{z})\right\vert d^{4}\widetilde{z}<\infty. \end{align*} \end{proof} The space of test functions $\mathbf{S}\left( \mathbb{Q}_{p}^{4}\right) $ is not invariant under the action of $\boldsymbol{f}\left( \partial ,\alpha\right) $. But if we replace $\mathbf{S}\left( \mathbb{Q}_{p ^{4}\right) $ by $\mathbf{\Phi}\left( \mathbb{Q}_{p}^{4}\right) $ then $\boldsymbol{f}\left( \partial,\alpha\right) \mathbf{\Phi}\left( \mathbb{Q}_{p}^{4}\right) =\mathbf{\Phi}\left( \mathbb{Q}_{p}^{4}\right) $. The verification of this fact involves the same ideas used in the verification of the corresponding assertion for the Taibleson operator, see e.g. \cite[Lemma 9.2.5]{A-K-S}. On the other hand, the mappin \ \begin{array} [c]{ccc \mathbf{\Phi}^{\prime}\left( \mathbb{Q}_{p}^{4}\right) & \rightarrow & \mathbf{\Phi^{\prime}}(\mathbb{Q}_{p}^{4})\\ & & \\ J & \rightarrow & \boldsymbol{f}\left( \partial,\alpha\right) J:=\mathcal{F ^{-1}\left[ \left\vert f^{\circ}\right\vert _{p}^{\alpha}\mathcal{F}\left[ J\right] \right] \end{array} \] is a homeomorphism. This is a consequence of the fact that the ma \ \begin{array} [c]{lll \mathbf{\Psi}\left( \mathbb{Q}_{p}^{4}\right) & \rightarrow & \mathbf{\Psi }\left( \mathbb{Q}_{p}^{4}\right) \\ & & \\ \varphi & \rightarrow & \left\vert pf^{\ast}\right\vert _{p}^{\alpha}\varphi \end{array} \] is a homeomorphism. \begin{lemma} \label{lemma7}The following formulas hold: (i) $\left\langle \boldsymbol{f}\left( \partial,\alpha\right) J,\varphi \right\rangle =\left\langle J,\boldsymbol{f}\left( \partial,\alpha\right) \varphi\right\rangle $ \ for any $J\in\mathbf{\Phi^{\prime}}(\mathbb{Q _{p}^{4})$ and $\varphi\in\mathbf{\Phi}(\mathbb{Q}_{p}^{4})$\thinspace; (ii) $\boldsymbol{f}\left( \partial,\alpha\right) J=K_{-\alpha}\ast J$ \ for any $J\in\mathbf{\Phi^{\prime}}(\mathbb{Q}_{p}^{4})$. \end{lemma} \begin{proof} (i) The formula follows from the following calculation \begin{multline*} \left\langle \boldsymbol{f}\left( \partial,\alpha\right) J,\varphi \right\rangle =\left\langle \mathcal{F}^{-1}\left[ \left\vert f^{\circ }\right\vert _{p}^{\alpha}\mathcal{F}\left[ J\right] \right] ,\varphi \right\rangle =\left\langle J,\mathcal{F}\left[ \left\vert f^{\circ }\right\vert _{p}^{\alpha}\mathcal{F}^{-1}\left[ \varphi\right] \right] \right\rangle \\ =\left\langle J,\mathcal{F}\left[ \left\vert f^{\circ}\left( -\xi\right) \right\vert _{p}^{\alpha}\mathcal{F}\left[ \varphi\right] \left( -\xi\right) \right] \right\rangle =\left\langle J,\mathcal{F}^{-1}\left[ \left\vert f^{\circ}\left( \xi\right) \right\vert _{p}^{\alpha \mathcal{F}\left[ \varphi\right] \left( \xi\right) \right] \right\rangle \\ =\left\langle J,\boldsymbol{f}\left( \partial,\alpha\right) \varphi \right\rangle . \end{multline*} (ii) The formula follows from the fact that $\left\vert f^{\circ}\right\vert _{p}^{\alpha}\mathcal{F}\left[ J\right] \in\mathbf{\Psi^{\prime} (\mathbb{Q}_{p}^{4})$ by using Proposition \ref{fourietransformtheorem}. \end{proof} \begin{definition} \label{Classical_sol}Consider $\boldsymbol{f}\left( \partial,\alpha\right) :$ $\mathbf{\Phi}(\mathbb{Q}_{p}^{4})\rightarrow\mathbf{\Phi}(\mathbb{Q _{p}^{4})$, and the equatio \begin{equation} \boldsymbol{f}\left( \partial,\alpha\right) u=\varphi\text{, }\quad \varphi\in\mathbf{\Phi}(\mathbb{Q}_{p}^{4})\text{.} \label{Equation \end{equation} A classical \ solution of (\ref{Equation}) is a function $u$ belonging to the domain of $\boldsymbol{f}\left( \partial,\alpha\right) $ which satisfies the equation. A fundamental solution of (\ref{Equation}) is a distribution $E_{\alpha}\in\mathbf{\Phi}^{\prime}\left( \mathbb{Q}_{p}^{4}\right) $ such that $u\left( x\right) =\left( E_{\alpha}\ast\varphi\right) \left( x\right) $ is a classical solution of (\ref{Equation}) for any $\varphi \in\mathbf{\Phi}(\mathbb{Q}_{p}^{4})$. \end{definition} \begin{lemma} \label{lemma8}The following two assertions are equivalent: (i) $E_{\alpha}\in\mathbf{\Phi}^{\prime}\left( \mathbb{Q}_{p}^{4}\right) $ is a fundamental solution of (\ref{Equation}); (ii) $\boldsymbol{f}\left( \partial,\alpha\right) E_{\alpha}=\delta$ in $\mathbf{\Phi}^{\prime}\left( \mathbb{Q}_{p}^{4}\right) $. \end{lemma} \begin{proof} (i) $\Leftrightarrow\boldsymbol{f}\left( \partial,\alpha\right) \left( E_{\alpha}\ast\varphi\right) =\varphi$ for any $\varphi\in\mathbf{\Phi }(\mathbb{Q}_{p}^{4})\Leftrightarrow\left\vert f^{\circ}\right\vert _{p}^{\alpha}\mathcal{F}\left[ E_{\alpha}\right] =1$ in $\mathbf{\Psi }^{\prime}\left( \mathbb{Q}_{p}^{4}\right) \Leftrightarrow\boldsymbol{f \left( \partial,\alpha\right) E_{\alpha}=\delta$ in $\mathbf{\Phi}^{\prime }\left( \mathbb{Q}_{p}^{4}\right) $. \end{proof} \begin{theorem} \label{mainB}(i) The functio \[ E_{\alpha}\left( x\right) =\left\{ \begin{array} [c]{lll \dfrac{1-p^{-\alpha}}{1-p^{\alpha-2}}|f(x)|_{p}^{\alpha-2} & \text{if} & \alpha\neq2\\ & & \\ -\dfrac{1-p^{-2}}{\ln p}\ln|f\left( x\right) |_{p} & \text{if} & \alpha=2 \end{array} \right. \] is a fundamental solution of (\ref{Equation}). (ii) Consider $\left\vert f\right\vert _{p}^{s}\in\mathbf{\Phi}^{\prime }\left( \mathbb{Q}_{p}^{4}\right) $, $s\in\mathbb{C}$. Then$\ \[ \boldsymbol{f}\left( \partial,1\right) \left\vert f\right\vert _{p ^{s+1}=\frac{\left( 1-p^{s+1}\right) \left( 1-p^{-s-2}\right) }{\left( 1-p^{-s-3}\right) \left( 1-p^{s}\right) }\left\vert f\right\vert _{p ^{s}\text{ in \ }\mathbf{\Phi}^{\prime}\left( \mathbb{Q}_{p}^{4}\right) . \] Here we are identifying $\left\vert f\right\vert _{p}^{s}$ with its meromorphic continuation. \end{theorem} \begin{proof} (i) By Lemma \ref{lemma8}, we have to show the existence of a distribution $E_{\alpha}$ in $\mathbf{\Phi}^{\prime}\left( \mathbb{Q}_{p}^{4}\right) $ satisfying $\boldsymbol{f}\left( \partial,\alpha\right) E_{\alpha}=\delta$, which is equivalent (by Lemma \ref{lemma7}\ (ii)) to solve $K_{-\alpha}\ast E_{\alpha}=\delta$. By Theorem \ref{mainA} this equation has unique solution $E_{\alpha}=K_{\alpha}$. Finally $u=E_{\alpha}\ast\varphi=\mathcal{F ^{-1}\left( \frac{\mathcal{F}\left( \varphi\right) }{\left\vert f^{\circ }\right\vert _{p}^{\alpha}}\right) \in\mathbf{\Phi}(\mathbb{Q}_{p}^{4})$ for $\varphi\in\mathbf{\Phi}(\mathbb{Q}_{p}^{4})$. (ii) Note that \[ \left\vert f\right\vert _{p}^{s+1}=\frac{1-p^{s+1}}{1-p^{-s-3}}K_{s+3}\text{ \ in \ }\mathbf{\Phi}^{\prime}\left( \mathbb{Q}_{p}^{4}\right) \text{ for }s\notin\left\{ -3+\alpha_{k}\right\} \cup\left\{ -1+\alpha_{k}\right\} . \] Then by Lemma \ref{lemma7}\ (ii) and Theorem \ref{mainA \begin{multline*} \boldsymbol{f}\left( \partial,1\right) \left\vert f\right\vert _{p ^{s+1}=K_{-1}\ast\left\vert f\right\vert _{p}^{s+1}=\left( \frac{1-p^{s+1 }{1-p^{-s-3}}\right) K_{-1}\ast K_{s+3}=\left( \frac{1-p^{s+1}}{1-p^{-s-3 }\right) K_{s+2}\\ =\frac{\left( 1-p^{s+1}\right) \left( 1-p^{-s-2}\right) }{\left( 1-p^{-s-3}\right) \left( 1-p^{s}\right) }\left\vert f\right\vert _{p ^{s}\text{ \end{multline*} in \ $\mathbf{\Phi}^{\prime}\left( \mathbb{Q}_{p}^{4}\right) $ for $s\notin\left\{ -1+\alpha_{k}\right\} \cup\left\{ -2+\alpha_{k}\right\} \cup\left\{ -3+\alpha_{k}\right\} \cup\left\{ \alpha_{k}\right\} $. The announced formula follows by analytic continuation, since the distributions $\boldsymbol{f}\left( \partial,1\right) \left\vert f\right\vert _{p}^{s+1}$ and $\frac{\left( 1-p^{s+1}\right) \left( 1-p^{-s-2}\right) }{\left( 1-p^{-s-3}\right) \left( 1-p^{s}\right) }\left\vert f\right\vert _{p}^{s}$ agree on an open and connected subset of the complex plane. \end{proof} \begin{remark} \label{nota_ultima}Similar results are valid for pseudodifferential operators attached to elliptic quadratic forms of dimension $2$. \end{remark}
\section{Introduction}\label{intro:sec} The set of exceptional slopes on a boundary component of a cusped hyperbolic manifold has generated a lot of interest in the literature; there are many restrictions on the set of exceptional slopes on a boundary component of a hyperbolic 3-manifold $M$ and its corresponding fillings, for example, no such $M$ has two distinct $S^3$ fillings \cite{GL} or more than 10 exceptional slopes \cite{Lack:setsize}. However, it is still not known if a hyperbolic knot exterior in $S^3$ can have a reducible filling, or if $M$ can have a pair of exceptional slopes $\alpha$ and $\beta$ corresponding to a lens space and toroidal space so that the \textit{distance} (minimal number of intersections) between $\alpha$ and $\beta$ is 4. Conjecturally, neither are possible, see \cite{Cabling} and \cite{3or4} respectively. Moreover, it is conjecturally the case that the figure-eight knot exterior is the unique one-cusped hyperbolic manifold with 10 exceptional slopes \cite{Gor'} and that the figure-eight knot exterior and its sister are the unique one-cusped hyperbolic manifolds having two exceptional slopes at distance 8, see \cite{problems}. In this article, by classifying the sets of exceptional slopes and the corresponding fillings for all manifolds obtained by surgery on the minimally twisted 5-chain link (see the rightmost link in Figure \ref{sequence:fig}), we provide experimental evidence that supports these conjectures: \begin{thm}\label{flash} If $M$ is a cusped hyperbolic manifold obtained by surgery on the minimally twisted 5-chain link then: \begin{itemize} \item $M$ is not the exterior of a knot with a reducible filling; \item $M$ does not have a lens space and toroidal filling at distance 4; \item If $M$ has 10 exceptional slopes then $M$ is the figure-8 knot exterior; \item If $M$ has exceptional slopes at distance 8 then $M$ is either the figure-8 knot exterior or the figure-8 knot exterior sister. \end{itemize} \end{thm} \subsection{The minimally twisted 5-chain link} A notable collection of chain links is described in \cite{Big}; they are the figure-8 knot, the Whitehead link, the 3-chain link, the 4-chain link with a half twist, and the minimally twisted 5-chain link. These links are shown in Figure \ref{sequence:fig}. We follow \cite{Big} and denote these chain links by 1CL, 2CL, 3CL, 4CL and 5CL, and their exteriors by $M_1$, $M_2$, $M_3$, $M_4$, $M_5$ respectively. \begin{figure}[h!] \begin{center} \includegraphics[width = 12 cm]{sequence.pdf} \end{center} \caption{The links 1CL, 2CL, 3CL, 4CL, 5CL in $S^3$ whose exteriors we denote by $M_1$, $M_2$, $M_3$, $M_4$, and $M_5$ respectively.} \label{sequence:fig} \end{figure} The significance of this sequence of links comes from the following facts: each $M_i$ is the (or conjecturally the) smallest volume hyperbolic 3-manifold with $i$ cusps, see \cite{Ago} and \cite{Yoshida}; more than $80\%$ of the cusped hyperbolic 3-manifolds from the Callahan-Hildebrand-Weeks census \cite{CaHiWe} are surgeries on 5CL (personal communication with Nathan Dunfield); 5CL relates to the program of enumerating exceptional pairs at maximal distance, in particular, all knots realising half-integral toroidal surgeries \cite{essential:tori}, many of the knots realising lens space surgeries \cite{Baker}, and all cusped hyperbolic 3-manifolds with distinct reducible and toroidal fillings at maximal distance \cite{Kang} are obtained by surgery on 5CL. It is easy to see that if $\partial M_n$ is equipped with the usual (meridian, longitude) homology basis then a $-1$ filling on any boundary component of $M_n$ results in $M_{n-1}$, and, as a result, that any manifold obtained by surgery on $(n-1)$CL is obtained by surgery on $n$CL. A classification of the exceptional surgeries on 3CL is given in \cite{Magic} together with a complete description of the set of exceptional slopes, and corresponding exceptional fillings, on the boundary components of all hyperbolic manifolds obtained by surgery on 3CL. A classification of exceptional surgeries on 5CL is found in \cite{Big}; in this article we completely describe the sets of exceptional slopes and corresponding exceptional fillings of any manifold obtained by surgery on 5CL. To simplify matters, we only consider the `new' manifolds obtained by surgery on 5CL; namely we only consider surgeries on 5CL and 4CL that are not obtained by surgery on 3CL. In particular, Theorems \ref{5CL_thm} and \ref{4CL_thm} provide explicit descriptions of the sets of exceptional slopes and corresponding fillings for all hyperbolic surgeries of 5CL not obtained by surgery on 3CL, and Theorem \ref{flash} will be an easy consequence. \subsection{Article structure} We start with Section \ref{basics} where we recall the classification from \cite{Big}. In order to do so, we begin by recalling and introducing terminology in Subsection \ref{introbasics}. The main results of this article are stated and proved in Section \ref{mr} and refer to the tables found in Section \ref{tables}. \subsection{Acknowledgements} The results of this article were obtained as a graduate student at the University of Pisa under the supervision of Carlo Petronio and Bruno Martelli. The current presentation of the results has benefited from discussions with Daniel Matignon and Mark Lackenby, and from email correspondences with Carlo Petronio, Bruno Martelli and Nathan Dunfield. \section{Background Terminology and Results}\label{basics} \subsection{Terminology}\label{introbasics} We begin with some general terminology, and we introduce the notion of an ``exceptional filling instruction" used in Theorems \ref{5CL_thm} and \ref{4CL_thm}. Fix an orientable compact 3-manifold $X$ with $\partial X$ consisting of tori: \begin{itemize} \item A \emph{slope} on a boundary component $\tau$ of $X$ is the isotopy class of a non-trivial unoriented loop on $\tau$; \item A \emph{filling instruction} $\alpha$ for $X$ is a set consisting of either a slope or the empty set for each component of $\partial X$; \item The \emph{filling} $X(\alpha)$ given by an instruction $\alpha$ is the manifold obtained by attaching one solid torus to $\partial X$ for each (non-empty) slope in $\alpha$, with the meridian of the solid torus attached to the slope. \end{itemize} We recall that if $M$ is a hyperbolic non-compact finite-volume 3-manifold then $M=\text{int}(X)$ with $\partial X$ consisting of tori, and that $\text{int}(X(\alpha))$ is hyperbolic for all but finitely many $\alpha$'s consisting of one slope and $\varnothing$'s \cite{Lack:setsize}. \begin{itemize} \item If the interior of $X$ is hyperbolic but the interior of $X(\alpha)$ is not, we say that $\alpha$ is an \emph{exceptional filling instruction} for $X$ and that $X(\alpha)$ is an \emph{exceptional filling} of $X$; \item We say that an exceptional filling instruction $\alpha$ on a hyperbolic $X$ is \emph{isolated} if $X(\beta)$ is hyperbolic for all $\beta$ properly contained in $\alpha$; for such an $\alpha$ we call $X(\alpha)$ an \emph{isolated exceptional filling} of $X$. \end{itemize} A \emph{surgery} on a link $L$ corresponds to a filling of the exterior of $L$, that is, to a filling of $M\backslash N(L)$ where $N(L)$ is an open regular neighbourhood of $L$. By a \emph{surgery instruction} for $L$ we mean a filling instruction on the exterior of $L$. We now recall some standard notation used in the description of the set of exceptional slopes on a fixed boundary component of a hyperbolic manifold, see for example \cite{Gor}. If $X$ is a hyperbolic 3-manifold with boundary consisting of tori and $\tau$ is a fixed boundary component of $\partial X$ then the set of exceptional slopes on $\tau$ is denoted by $E_{\tau}(X)$, and the cardinality of $E_{\tau}(X)$ by $e_{\tau}(X)$. To describe $E_{\tau}(M_5(\alpha))$ we introduce the following definition: \begin{definition} Let $\alpha$ be a filling instruction on a manifold $X$. We say that $\alpha$ \emph{factors through} a manifold $Y$ if there exists some filling instruction $\beta \subset \alpha$ such that $Y=X(\beta)$. \end{definition} Note that if $\alpha$ is exceptional for $X$ and factors through a hyperbolic $Y$ with $Y=X(\beta)$, then $\alpha \backslash \beta$ is exceptional for $Y$. \subsection{Exceptional fillings of $M_5$}\label{excep:fill:sec} Our description of the exceptional fillings of $M_5(\gamma)$ will employ the notation now discussed for Seifert manifolds with orientable base surface. Given integers $p_1,\dots, p_n, q_1,\dots, q_n$, with $p_i>1$ and $G$ an orientable surface with $k\geq0$ boundary components $b_1\dots b_k$, we let $\Sigma$ denote the surface obtained by removing $n$ open discs from $G$ and we denote by $b_{k+1}, \dots, b_{k+n}$ the $n$ newly introduced boundary circles. We fix an orientation on $\Sigma\times S^1$ and orient $\{\mu_i, \lambda_i\}=\{b_i\times \{\ast\}, \{\ast\}\times S^1\}$ so that $\mu_i$, $\lambda_i$ is a positive basis of $H_1(b_i \times S^1)$ with $b_i \times S^1$ oriented as $\partial (\Sigma\times S^1)$. We denote by $(G, (p_1, q_1), \dots (p_n, q_n))$, the manifold obtained by performing a Dehn filling on each $b_i \times S^1$ along $p_i\mu_i + q_i\lambda_i$ for $i>k$. In our case, $G$ will be either the disc $D$, the annulus $A$, or the sphere $S^2$. Given Seifert manifolds $X$ and $Y$ with orientable base surfaces with boundary as described above, and $B\in\textrm{GL}(2,\mathbb{Z})$ with $\det(B)=-1$, we define $X\bigcup_B Y$ unambiguously to be the quotient manifold $X\bigcup_f Y$ where $f:T\rightarrow U$ for $T$ and $U$ arbitrary boundary components of $X$ and $Y$ respectively, and $f$ acting on homology by $B$ with respect to the bases described above. The case $T, U\subset \partial X$, $T\neq U$ is also allowed and we write the quotient manifold as $X\big/_B$. For the purposes of Theorem \ref{5CL_thm} and \ref{4CL_thm}, we employ a flexible notation for Seifert manifolds. We will formally identify an $\varnothing$ slope with $\frac 00$ and allow parameters $(p_i, q_i)$ to be arbitrary elements of $\mathbb{Z}^2$. \subsection{Exceptional surgery instructions on 5CL}\label{result:recap} By ordering the components of 5CL cyclically, surgery instructions on 5CL can be naturally identified with $(\mathbb{Q}\cup\{\varnothing, \infty\})^5$. Given such an instruction $\alpha$, the symmetry group of $M_5(\alpha)$ has a natural action on the boundary components of $M_5(\alpha)$. This action induces an action on the filling instructions on $M_5(\alpha)$. Among the most significant actions arising we mention those coming from the symmetry group of $M_5$, see (\ref{first:eqn})-(\ref{3rd:eqn}) below, a symmetry of $M_4$ which may be deduced from the Fenn-Rourke blow-down move on 5CL, see (\ref{blow:eqn}), and from the amphichirality of the figure-8 knot $M_5(-1,-2,-2,-2,\varnothing)$, see (\ref{figure8:eqn}); see \cite{Big} for full details. \begin{align} (\alpha_1, \alpha_2,\alpha_3,\alpha_4,\alpha_5) & \longmapsto (\alpha_5, \alpha_1, \alpha_2, \alpha_3, \alpha_4) \label{first:eqn} \\ (\alpha_1, \alpha_2, \alpha_3, \alpha_4, \alpha_5) & \longmapsto (\alpha_5, \alpha_4, \alpha_3, \alpha_2, \alpha_1) \label{second:eqn} \\ (\alpha_1, \alpha_2,\alpha_3,\alpha_4,\alpha_5) & \longmapsto (\alpha_2^{-1}, \alpha_1^{-1}, 1-\alpha_3, (1-\alpha_4)^{-1}, 1-\alpha_5)\label{3rd:eqn}\\ (-1, \alpha_1, \alpha_2, \alpha_3, \alpha_4) & \longmapsto (\alpha_1, -1, \alpha_2-1, \alpha_3, \alpha_4+1)\label{blow:eqn}\\ \left(-1, -2, -2, -2, \alpha \right) & \longmapsto \left(-1, -2, -2, -2, -\alpha - 6 \right)\label{figure8:eqn} \end{align} Since the action of the maps (\ref{first:eqn}) and (\ref{second:eqn}) is very easily understandable while that of (\ref{3rd:eqn})-(\ref{figure8:eqn}) is more involved, we introduce the symbol $[\alpha]$ for the equivalence class of a filling instruction $\alpha$ under (\ref{first:eqn}) and (\ref{second:eqn}) only, and the symbol $[\![\alpha]\!]$ for the equivalence class of $\alpha$ under (\ref{first:eqn})-(\ref{figure8:eqn}). For a filling instruction $\alpha$ on $M_5$ we will often simplify notation by omitting empty slopes but leaving the subscripts on non-empty slopes. For example, $((-1)_1, (-1)_3)$ corresponds to the filling instruction $(\varnothing, \mu_1-\lambda_1, \varnothing, \mu_3-\lambda_3, \varnothing)$ with $(\mu_i, \lambda_i)$ the (meridian, longitude) basis of the homology of the $i^{\text{th}}$ cusp. Note that for any $\frac pq\in\mathbb{Q}\cup \{\infty\}$ and $i\neq j$ one has $[((\frac pq)_i)]=[((\frac pq)_j)]$, so the fillings $M_5(\frac pq)$ are defined without ambiguity. Our convention will be that a filling instruction $(\frac pq, \frac rs, \frac uv, \frac xy)$ on $M_5$ with four non-empty slopes and no subscripts represents $({\frac pq}, {\frac rs}, {\frac uv}, {\frac xy}, \varnothing)$. We now state the main result of \cite{Big}. \begin{thm} \label{main:teo} Every exceptional filling instruction on $M_5$ is equivalent up to a composition of the symmetries \emph{(\ref{first:eqn})-(\ref{figure8:eqn})} to a filling instruction containing one of: \begin{align*} \infty, (-1, -2, -2, -1), & (-1, -2, -3, -2, -4), (-1, -2, -2, -3, -5),\\ (-1, -3, -2, -2, -3),& \left(-2, -\tfrac 12, 3, 3, -\tfrac 12\right), (-2, -2, -2, -2, -2). \end{align*} \end{thm} Descriptions of the corresponding exceptional fillings may be found in \cite{Big}. We remark that the second, third, and fourth filling instructions factor through $M_3$, and we assign the names $m_1$, $m_2$, and $m_3$ to the manifolds $M_5(-1, -3, -2, -2, -3)$, $M_5\left(-2, -\tfrac 12, 3, 3, -\tfrac 12\right)$, and $M_5(-2, -2, -2, -2, -2)$ respectively. \section{Main Results}\label{mr} We now precisely describe the classification of exceptional slopes of every hyperbolic surgery on 5CL by describing $E_{\tau}(M_5(\alpha))$ for every $\alpha$ not factoring through $M_3$. We split the result according to whether or not $\alpha$ factors through $M_4$. \begin{thm}\label{5CL_thm} Let $\gamma$ be a hyperbolic filling instruction on $M_5$ containing at least one $\varnothing$ and not factoring through $M_4$, and let $\tau$ be a boundary component of $\partial M_5(\gamma)$; then either $e_{\tau}(M_5(\gamma))=3$ and, with respect to the basis induced from $M_5$, we have $E_{\tau}(M_5(\gamma))=\{0,1,\infty\}$ and \begin{gather} M_5(\tfrac ab, \tfrac cd, \tfrac ef, \tfrac gh)(\infty)= \seifdue{D}a{-b}dc \bigu0110 \seifdue{D}feg{-h},\label{5CLinf} \\ M_5(\tfrac ab, \tfrac cd, \tfrac ef, \tfrac gh)(1)= \seifdue{D}{a-b}{b}ef \bigu0110 \seifdue{D}{g-h}hcd, \label{5CL1} \\ M_5(\tfrac ab, \tfrac cd, \tfrac ef, \tfrac gh)(0)= \seifdue{D}{b}{b-a}{h}{-g} \bigu0110 \seifdue{D}{c-d}c{e-f}{f}, \label{5CL0} \end{gather} or $4\leqslant e_{\tau}(M_5(\gamma)) \leqslant 5$ and \begin{itemize} \item $M_5(\gamma)\cong M_5(\alpha)$ for some $\alpha$ in Tables \emph{\ref{tb1}}-\emph{\ref{tb4}} in Section \ref{tables}, and with respect to the basis induced from $M_5$ we have \begin{equation*} E_{\tau}(M_5(\alpha))= \begin{cases} \{\beta_1,\beta_2, 0,1,\infty\} \mbox{ if $\alpha$ is found in Table \emph{\ref{tb1}}};\\ \{\beta, 0,1,\infty\} \mbox{ otherwise}, \end{cases} \end{equation*} where $\beta, \beta_1,\beta_2$ are found in Tables \emph{\ref{tb1}}-\emph{\ref{tb4}}. \item Identities \emph{(\ref{5CLinf})-(\ref{5CL0})} hold and $M_5(\alpha)(\beta_i)$, $M_5(\alpha)(\beta)$ are shown in Tables \emph{\ref{tb1}}-\emph{\ref{tb4}}. \end{itemize} \end{thm} Ordering the components of $M_4$ cyclically, identifying filling instructions on $M_4$ with $(\mathbb{Q}\cup\{\varnothing,\infty\})^4$ and adopting the convention that a filling instruction $(\frac pq, \frac rs, \frac uv)$ on $M_4$ with three slopes and no subscripts represents $({\frac pq}, {\frac rs}, {\frac uv}, \varnothing)$, the corresponding result for $M_4$ is: \begin{thm}\label{4CL_thm} Let $\gamma$ be a hyperbolic filling instruction on $M_4$ containing at least one $\varnothing$ and not factoring through $M_3$, and let $\tau$ a boundary component of $\partial M_4(\gamma)$; then either $e_{\tau}(M_4(\gamma))=4$ and, with respect to the basis induced from $M_4$, we have $E_{\tau}(M_4(\gamma))=\{0,1,2,\infty\}$ and \begin{gather} M_4(\tfrac ab, \tfrac cd, \tfrac ef)(\infty)= \seiftre{S^2}abd{-c}ef, \label{4CLinf} \\ M_4(\tfrac ab, \tfrac cd, \tfrac ef)(2)= \seifdue{D}{a-b}b{e-f}f \bigu0110 \seifdue{D}cd2{-1}, \label{4CL2} \\ M_4(\tfrac ab, \tfrac cd, \tfrac ef)(1)= \seiftre{S^2}{a-2b}b{c-d}c{e-2f}f, \label{4CL1} \\ M_4(\tfrac ab, \tfrac cd, \tfrac ef)(0)= \seifdue{D}f{-e}b{2b-a} \bigu0110 \seifdue{D}21{c-2d}d, \label{4CL0} \end{gather} or $5\leqslant e_\tau(M_4(\gamma))\leqslant 6$ and \begin{itemize} \item $M_4(\gamma)\cong M_4(\alpha)$ for some $\alpha$ in Tables \emph{\ref{tb5}}-\emph{\ref{tb6}}, in Section \ref{tables}, and with respect to the basis induced from $M_4$ we have \begin{equation*} E_{\tau}(M_4(\alpha))= \begin{cases} \{\beta_1,\beta_2, 0,1,2,\infty\} \mbox{ if $\alpha$ is found in Table \emph{\ref{tb5}}};\\ \{\beta, 0,1,2,\infty\} \mbox{ otherwise}, \end{cases} \end{equation*} where $\beta, \beta_1,\beta_2$ are found in Tables \emph{\ref{tb5}} and \emph{\ref{tb6}}. \item Identities \emph{(\ref{4CLinf})-(\ref{4CL0})} hold, and the $M_5(\alpha)(\beta_i)$, $M_4(\alpha)(\beta)$ are shown in Tables \emph{\ref{tb5}} and \emph{\ref{tb6}}. \end{itemize} \end{thm} While proving Theorems \ref{5CL_thm} and \ref{4CL_thm} we will often see identities between fillings of the $M_i$. When a homeomorphism $M_5(\alpha)\simeq M_5(\beta)$ follows as a consequence of maps (\ref{first:eqn})-(\ref{figure8:eqn}) we will write $M_5(\alpha)\sim M_5(\beta)$. In addition to realising homeomorphisms between fillings of $M_5$, we will often use the Rolfsen twist (see \cite{Rolfsen}) to obtain homeomorphisms between fillings of the $M_i$; when $M_5(\alpha)\simeq M_k(\beta)$ is induced from a (composition of) Rolfsen twist(s), we will write $M_5(\alpha)\simeq M_k(\beta)$. \vspace{.5cm} \textsc{Proof of Theorem \ref{5CL_thm}.} By (\ref{first:eqn})-(\ref{figure8:eqn}), we have \begin{align} [\![(1)]\!]& = [(1)]\sqcup[(\infty)]\sqcup[(0)]; \label{1action} \\ [\![(-1)]\!]& = [(-1)]\sqcup[(\tfrac12)]\sqcup[(2)]; \label{-1action} \\ [\![(-2)]\!]& = [(-2)]\sqcup [(-\tfrac12)]\sqcup [(\tfrac13)]\sqcup [(\tfrac23)]\sqcup [(\tfrac32)]\sqcup [(3)] \label{-2action}. \end{align} Theorem \ref{main:teo} implies that all slopes in $[\![(1)]\!]$ are exceptional, moreover, any instruction containing a slope in $[\![(-1)]\!]$ factors through $M_4$. So for $\tau$ a boundary component of a hyperbolic $M_5(\gamma)$ we have $\{0,1,\infty\}\subseteq E_{\tau}(M_5(\gamma))$ and $e_{\tau}(M_5(\gamma)) \geq 3$; when $\gamma$ does not factor through $M_4$ no slope in $[\![(-1)]\!]$ is contained in $\gamma$. Identity (\ref{5CLinf}) is established in \cite{Big}, and Identities (\ref{5CL1}) and (\ref{5CL0}) are established using (\ref{first:eqn})-(\ref{3rd:eqn}) and (\ref{5CLinf}). We will now describe up to (\ref{first:eqn})-(\ref{3rd:eqn}) all hyperbolic $\gamma$'s not factoring through $M_4$ with $e_{\tau}(M_5(\gamma))>3$. For such $\gamma$, using (\ref{first:eqn})-(\ref{second:eqn}), we may assume that $\tau$ corresponds to the neighbourhood of the fourth chain component of 5CL. So we suppose that $\gamma_0, \gamma_1, \gamma_2, \gamma_3\in \mathbb{Q}\cup\{\varnothing\}$, with $\gamma=(\gamma_0, \gamma_1, \gamma_2, \gamma_3)$ a hyperbolic filling instruction on $M_5$ that does not factor through $M_4$. If $\beta$ is an exceptional slope on $M_5(\gamma)$ then, up to (\ref{first:eqn})-(\ref{3rd:eqn}), $(\gamma_0,\gamma_1, \gamma_2,\gamma_3, \beta)$ contains one of the seven isolated exceptional filling instruction from Theorem \ref{main:teo}. If $\beta\neq 0,1,\infty$ then any such isolated exceptional filling instruction contains at most one slope in $[\![(-1)]\!]$, and contains two slopes in $[\![(-2)]\!]$ (see Theorem \ref{main:teo} and (\ref{-2action})). Thus at least one of the slopes in $\gamma$ belongs to $[\![(-2)]\!]$. It is a routine consequence of maps (\ref{first:eqn})-(\ref{3rd:eqn}) that we may assume without loss of generality that $\alpha_0=-2$ and that $\tau$ remains the boundary corresponding to the neighbourhood of the 4th chain component of 5CL. So for $\gamma$ hyperbolic, not factoring through $M_4$, we have $[\![\gamma]\!]=[\![\alpha]\!]$ where $\alpha_i \not \in\{-1, \frac 12, 2\}$ and $\alpha_0=-2$. We define $(\alpha, \beta)$ to be $(\alpha_0, \alpha_1, \alpha_2, \alpha_3, \beta)$. If $\beta$ is an exceptional slope of $M_5(\alpha)$ then $(\alpha, \beta)$ contains an isolated exceptional filling instruction by Theorem \ref{main:teo}. Moreover, using Theorem \ref{main:teo}, if $(\alpha, \beta)$ contains an isolated exceptional filling instruction then either $\beta$ is a slope in $[\![(1)]\!]$, or $\beta$ is a slope in $[\![(-1)]\!]$ and $(\alpha, \beta)$ factors through $M_3$, or $M_5(\alpha, \beta)\in \{m_1, m_2, m_3\}$. As we are assuming that $\alpha$ does not factor through $M_4$, if $(\alpha, \beta)$ factors through $M_3$ then we require $\beta\in\{-1, \frac 12, 2\}$. We define the following sets of $(\alpha, \beta)$'s; \begin{itemize} \item We define $l$ to be the set of exceptional $(\alpha, \beta)$'s such that $M_5(\alpha, \beta)$ is in $\{m_1, m_2, m_3\}$; \item We define $l_{-1}$ to be the set of exceptional $(\alpha, \beta)$'s factoring through $M_3$ with $\beta=-1$; \item We define $l_{\frac 12}$ to be the set of exceptional $(\alpha, \beta)$'s factoring through $M_3$ with $\beta=\frac 12$; \item We define $l_{2}$ to be the set of exceptional $(\alpha, \beta)$'s factoring through $M_3$ with $\beta=2$; \end{itemize} We know that, every $\gamma$ with $M_5(\gamma)$ a hyperbolic manifold, $\gamma$ not factoring through $M_4$, and $e_{\tau}(M_5(\gamma))>3$ has $[\![\gamma]\!]=[\![\alpha]\!]=[\![(\alpha_0,\alpha_1,\alpha_2,\alpha_3)]\!]$ where $\alpha_0=-2$. The set $l_{-1}\cup l_{\frac 12}\cup l_2$ is the set of all exceptional filling instructions $(\alpha, \beta)$ factoring through $M_3$ with $M_5(\alpha)$ hyperbolic not factoring through $M_4$ and $\alpha_0=-2$, and the set $l$ is the set of all exceptional filling instructions $(\alpha, \beta)$ not factoring through $M_3$ with $\alpha_0=-2$. So, up to (\ref{first:eqn})-(\ref{3rd:eqn}) every hyperbolic filling instruction $\gamma$ not factoring through $M_4$ with $e_{\tau}(M_5(\gamma))>3$ is equivalent to a filling instruction $\alpha$ of the form $(-2, \alpha_1, \alpha_2, \alpha_3)$, every $(-2, \alpha_1, \alpha_2, \alpha_3)$ with $e_{\tau}(M_5(-2, \alpha_1, \alpha_2, \alpha_3)>3$ and $(-2, \alpha_1, \alpha_2, \alpha_3)$ not factoring through $M_4$ is some $\alpha$ in $l\cup l_{-1}\cup l_{\frac 12}\cup l_2$, and every exceptional slope on $M_5(-2, \alpha_1, \alpha_2, \alpha_3)$ not in $\{0, 1, \infty\}$ is some $\beta$ in $l\cup l_{-1}\cup l_{\frac 12}\cup l_2$. Now we can easily describe, up to $\sim$, all such $\alpha$ with more than 3 exceptional slopes. Let $p:l\cup l_{-1}\cup l_{\frac 12}\cup l_{2}\rightarrow A$ be defined by $(\alpha, \beta) \mapsto \alpha$. For $\alpha$ in $p(l\cup l_{-1}\cup l_{\frac 12}\cup l_{2})$ define $B_{\alpha}$ to be the set of all $\beta$'s such that $(\alpha, \beta)$ is contained in $l\cup l_{-1}\cup l_{\frac 12}\cup l_2$. It is clear that $E_{\tau}\big(M_5(\alpha)\big)= \{0, 1, \infty\}\cup B_{\alpha}$ and that $\{\big(M_5(\alpha), E_{\tau}(M_5(\alpha))\big)\}_{\alpha}$ is a complete list of all $\big(M_5(\alpha), E_{\tau}(M_5(\alpha))\big)$ pairs with $\alpha= (-2, \alpha_1, \alpha_2, \alpha_3)$, $M_5(\alpha)$ hyperbolic and $e_{\tau}\big(M_5(\alpha)\big)>3$. We now explicitly construct the sets $l, l_{-1}, l_{\frac 12}, l_2$. \textbf{Construction of the set $l$:} We can see using maps (\ref{first:eqn})-(\ref{figure8:eqn}) that every $(\alpha, \beta)$ with $M_5(\alpha, \beta)=m_i$ with $\alpha_0=-2$ and $\alpha$ hyperbolic not factoring through $M_4$ is contained in the following set \begin{equation*} \begin{array}{l} l=\Big\{\big( -2, \tfrac 14, \tfrac 32, \tfrac 43, \tfrac 12 \big) , \big(-2, -\tfrac 12, 3, 3, -\tfrac 12\big), \big(-2, \tfrac 13, 3, \tfrac 13, -2\big), \big(-2, -2, \tfrac 13, 3, \tfrac 13\big),\\ \big(-2, -\tfrac 12, -2, \tfrac 32, \tfrac 32\big), \big(-2, \tfrac 32, \tfrac 32, -2, -\tfrac 12\big), \big(-2, -2, -2, -2, -2\big), \big(-2, \tfrac 13, \tfrac 32, \tfrac 32, \tfrac 13\big)\Big\}. \end{array} \end{equation*} Before constructing the sets $l_{-1}, l_{\frac 12}, l_2$ we make some initial remarks. It is easy to see that if a surgery instruction on 5CL contains $((-1)_0, (-1)_2)$, then it factors through $M_3$. Up to the action of maps (\ref{first:eqn})-(\ref{figure8:eqn}) the surgery instruction $((-1)_0, (-1)_2)$ is equivalent to the following surgery instructions: \begin{align*} \Big\{ [((-1)_0, (-2)_1)], [((\tfrac 12)_0, (\tfrac 23)_2)], [((\tfrac 12)_0, 3_1)], [((\tfrac 12)_0, (\tfrac 32)_1)], [((\tfrac 23)_0, 2_1)], \\ [((\tfrac12)_0, (\tfrac13)_2) ], [(2_0, (-\tfrac 12)_2)], [((\tfrac 13)_0, 2_1)], [((-1)_0, 3_2)], [((-1)_0,(\tfrac 32)_2)], \\ [(2_0, (-2)_2)], [((-1)_0, (-\tfrac 12)_1)], [((-1)_0, (-1)_2)], [((-1)_0, (\tfrac12)_2)], \\ [((\tfrac12)_0, (2)_2)], [((-1)_0, (2)_1)], [((-1)_0, (\tfrac12)_1)], [((2)_0, (2)_2)]\Big\} \end{align*} We will now see that the requirements of $\alpha$ being hyperbolic and not factoring through $M_4$ allow us to completely construct $l_{-1}, l_{\frac 12}, l_2$. \textbf{Construction of the set $l_{-1}$:} In this case, for $\alpha=(-2, \alpha_1, \alpha_2, \alpha_3)$, we have $M_5(\alpha)(-1)\simeq M_3(\alpha_1+1, \alpha_2, \alpha_3+2)$, and $(\alpha, \beta)$ is an exceptional filling of $M_5$ if and only if $(\alpha_1+1, \alpha_2, \alpha_3+2)$ contains an isolated exceptional filling instruction on $M_3$; i.e.\ using the classification in \cite{Magic} we can see that $-1$ is an exceptional slope on $M_5(\alpha)$ if $(\alpha_1+1, \alpha_2, \alpha_3+2)$ contains an isolated instruction in \begin{equation*} \begin{array}{c} \Big\{0,1,2,3,\infty,(-1,-1), (4,\tfrac12), (\tfrac32,\tfrac52),(5,5,\tfrac12), (4,4,\tfrac23), (4,\tfrac32,\tfrac32),(4,\tfrac13,-1),(\tfrac83,\tfrac32,\tfrac32),\\ (\tfrac52,\tfrac52,\tfrac43), (\tfrac52,\tfrac53,\tfrac53), (\tfrac73,\tfrac73,\tfrac32), (-1,-2,-2), (-1,-2,-3),(-1,-2,-4), (-1,-2,-5),\\ (-1,-3,-3),(-2,-2,-2) \Big\}, \end{array} \end{equation*} For $\alpha$ non-exceptional and not factoring through $M_4$, no slope in $\alpha$ is in [(1)] or $[(-1)]$; that is $\alpha_1+1\not\in\{\infty, 2, 1, 0, \frac 32, 3\}$, $\alpha_2 \not \in \{\infty, 1, 0, -1, \frac 12, 2\}$, $\alpha_3+2\not\in \{\infty, 3, 2, 1, \frac 52, 4\}$. With these conditions it is easy to see that $\beta$ is an exceptional slope on a hyperbolic $M_5(\alpha)$ if and only if one of the following holds: \begin{itemize} \item $\alpha_2 =3$ \\ \item $\alpha_3+2=0$ \\ \item $(\alpha_1+1, \alpha_3+2)$ belongs to $\{(-1, -1), (\frac 52, \frac 32), (4, \frac 12)\}$ \\ \item $(\alpha_2, \alpha_1 +1)$ belongs to $\{(\frac 32, \frac 52), (4, \frac 12)\}$ \\ \item $(\alpha_2, \alpha_3+2)$ belongs to $\{(\frac 52, \frac 32), (4, \frac 12)\}$ \\ \item $(\alpha_1 +1, \alpha_2, \alpha_3 +2)$ belongs to \begin{gather*} \Big\{(5, 5, \tfrac 12), (\tfrac 12, 5, 5),(4, 4, \tfrac 23), (4, \tfrac 32, \tfrac 32), (4, \tfrac13, -1), (\tfrac 13, 4, -1), (-1, 4, \tfrac 13), (\tfrac 83, \tfrac 32, \tfrac 32), \\ (\tfrac 52, \tfrac 52, \tfrac 43), (\tfrac 52, \tfrac 53, \tfrac 53), (\tfrac 53, \tfrac 52, \tfrac 53), (\tfrac 73, \tfrac 73, \tfrac 32), (\tfrac 73, \tfrac 32, \tfrac 73), (-1, -2, -2), (-2, -2, -1),\\ (-1, -2, -3), (-3, -2, -1), (-2, -3, -1), (-1, -3, -2), (-1, -2, -4),\\ (-4, -2, -1), (-1, -4, -2), (-2, -4, -1), (-1, -2, -5), (-5, -2, -1), \\ (-1, -5, -2), (-2, -5, -1), (-1, -3, -3), (-3, -3, -1), (-2, -2, -2))\Big\}. \end{gather*} \end{itemize} Thus, the set of all $(-2, \alpha_1, \alpha_2, \alpha_3, -1)$ constructed in the above analysis is the set $l_{-1}$. \textbf{Construction of the set $l_{\frac 12}$:} Reasoning as in the case $\beta=-1$ we see $(\alpha, \beta)$ factors through $M_3$ when one of $\alpha_1, \alpha_2$ is in $\{\frac 13, \frac 23\}$ or $\alpha_3 \in \{3, \frac 32\}$ (see $[\![(-1_0, -1_2)]\!]$ shown above). We examine each of these 6 cases individually and enumerate all $(-2, \alpha_1, \alpha_2, \alpha_3, \frac 12)$ with the desired properties. If $\alpha_3 = 3$ then \[M_5\big(-2, \alpha_1, \alpha_2, 3, \tfrac 12\big)\sim M_5\big(-1, 3, \alpha_2^{-1}, \alpha_1^{-1}, -2\big) \simeq M_3\big(5, \alpha_2^{-1}, \alpha_1^{-1}+1\big).\] The result is that $\frac 12$ is an exceptional slope on $M_5\big(-2, \alpha_1, \alpha_2, \alpha_3\big)$ if and only if one of the following holds; $\alpha_1^{-1} =3$, $\big(\alpha_2^{-1}, \alpha_1^{-1}+1\big)=\big(\frac 32, \frac 52\big)$, $(4, \frac 12)$, $(5, \frac 12)$. If $\alpha_3 = \frac 32$ then \begin{gather*} M_5\big(-2, \alpha_1, \alpha_2, \tfrac 32, \tfrac 12\big)\sim M_5\big(-2, -2, (1-\alpha_2)^{-1}, 1-\alpha_1^{-1}, -1\big) \\ \simeq M_3\big(-1, (1-\alpha_2)^{-1}, 3-\alpha_1^{-1} \big). \end{gather*} We conclude that $\frac 12$ is an exceptional slope on $M_5\big(-2, \alpha_1, \alpha_2, \alpha_3\big)$ if and only if one of the following holds; $(1-\alpha_2)^{-1}=3$, $3-\alpha_1^{-1} =0, -1$, $((1-\alpha_2)^{-1}, 3-\alpha_1^{-1} \big)=(\frac 52, \frac 32)$, $(4, \frac 12)$, $(-2, -2)$, $(-2, -3)$, $(-3, -2)$, $(-2, -4)$, $(-2, -5)$, $(-5, -2)$, $(4, \frac 13)$. If $\alpha_1 =\frac 13$ then \begin{gather*} M_5(\alpha)(\beta)\sim M_5\big((1-\alpha_3)^{-1}, -2, (1-\alpha_2)^{-1}, -2, -1\big)\\ \simeq M_3\big(2+(1-\alpha_3)^{-1},-2, 1+(1-\alpha_2)^{-1}\big). \end{gather*} The result is that $\frac 12$ is an exceptional slope on $M_5\big(-2, \alpha_1, \alpha_2, \alpha_3\big)$ if and only if one of the following holds; $1+(1-\alpha_2)^{-1}=3$, $2+(1-\alpha_3)^{-1}=0$, $(2+(1-\alpha_3)^{-1},1+(1-\alpha_2)^{-1})=(-1, -1)$, $(\frac 12, 4)$, $(\frac 32, \frac 52)$. If $\alpha_1=\frac 13$ then \[M_5\big(-2, \alpha_1, \alpha_2, \alpha_3, \tfrac 12\big)\sim M_5\big(\tfrac 13, -1, -2, (1-\alpha_1)^{-1}, \alpha_3\big) \simeq M_3\big(\tfrac 73, 1+(1-\alpha_1)^{-1}, \alpha_3\big).\] We see that $\frac 12$ is an exceptional slope on $M_5\big(-2, \alpha_1, \alpha_2, \alpha_3\big)$ if and only if one of the following holds; $\alpha_3=3$, $(\alpha_3, 1+(1-\alpha_1)^{-1})$=$(\frac 32, \frac 52)$, $(\frac 32, \frac 73)$. If $\alpha_1 =\frac 23$ then \begin{gather*} M_5(\alpha)(\beta)\sim M_5\big(\tfrac 23, (1-\alpha_2^{-1})^{-1}, -1, -2, (1-\alpha_3^{-1})^{-1}\big)\\ \simeq M_3\big(\tfrac 23, 2+(1-\alpha_2^{-1})^{-1}, 1+(1-\alpha_3^{-1})^{-1}\big). \end{gather*} We find that $\frac 12$ is an exceptional slope on $M_5\big(-2, \alpha_1, \alpha_2, \alpha_3\big)$ if and only if one of the following holds; $2+(1-\alpha_2^{-1})^{-1}=0$, $(2+(1-\alpha_2^{-1})^{-1}, 1+(1-\alpha_3^{-1})^{-1})=(-1, -1)$, $(\frac 32, \frac 52)$. If $\alpha_2 =\frac 23$ then \begin{gather*} M_5(\alpha)(\beta)\sim M_5\big(\tfrac 23, -2, -1, (1-\alpha_1^{-1})^{-1}, (1-\alpha_3^{-1})^{-1}\big) \\ \simeq M_3\big(\tfrac 53, 2+(1-\alpha_1^{-1})^{-1}, (1-\alpha_3^{-1})^{-1}\big). \end{gather*} The result is that $\frac 12$ is an exceptional slope on $M_5\big(-2, \alpha_1, \alpha_2, \alpha_3\big)$ if and only if one of the following holds; $(1-\alpha_3^{-1})^{-1}=3$, $(2+(1-\alpha_1^{-1})^{-1}, (1-\alpha_3^{-1})^{-1})=(\frac 32, \frac 52)$, $(\frac 12, 4)$, $(\frac 53, \frac 52)$. Thus, the set of all $(-2, \alpha_1, \alpha_2, \alpha_3, \frac 12)$ constructed in the above analysis is the set $l_{\frac 12}$. \textbf{Construction of the set $l_{2}$:} For $(\alpha, \beta)=(-2, \alpha_1, \alpha_2, \alpha_3, 2)$ to factor through $M_3$ we need one of $\alpha_1, \alpha_2$ to be in $\{-2, -\frac 12\}$ or $\alpha_3 \in \{\frac 13, \frac 23\}$ (see $[\![(-1_0, -1_2)]\!]$ shown above). We construct all $(\alpha, \beta)$ for each of these 6 cases individually. If $\alpha_1 = -2$ then \[M_5(\alpha)(2) \sim M_5\big(\tfrac 32, 1-\alpha_2^{-1}, (1-\alpha_3)^{-1}, -2, -1\big) \simeq M_3\big(\tfrac 72, 1-\alpha_2^{-1}, 1+(1-\alpha_3)^{-1}\big).\] When no $\alpha_1$, $\alpha_2$, $\alpha_3$ is a slope in $[\![(1)]\!], [\![(-1)]\!]$ we have $1-\alpha_2^{-1}\not\in\{\infty, 0, 1, 2, \frac 12, -1\}$ and $1+(1-\alpha_3)^{-1}\not\in\{2, \infty, 1, \frac 32, 0, 3\}$. From \cite{Magic} we see that 2 is exceptional if and only if one of the following holds; $1-\alpha_2^{-1}=3$, $(1-\alpha_2^{-1}, 1+(1-\alpha_3)^{-1})=(\frac 32, \frac 52)$, $(4, \frac 12)$. If $\alpha_2 = -2$ then \[M_5(\alpha)(\beta)\sim M_5\big(-1, -2, \tfrac 13, 1-\alpha_1^{-1}, 1-\alpha_3^{-1}\big) \simeq M_3\big(\tfrac 43, 1-\alpha_1^{-1}, 3-\alpha_1^{-1}\big).\] Using \cite{Magic} we get that $2$ is exceptional on $M_5(\alpha)$ if and only if one of the following holds; $1-\alpha_1^{-1}=3$, $3-\alpha_3^{-1}=0$, $(1-\alpha_1^{-1}, 3-\alpha_3^{-1})=(4, \frac 12)$, $(\frac 52, \frac 32)$. If $\alpha_1 =-\frac 12$ then \[ M_5(\alpha)(\beta)\sim M_5\big((\alpha_3-1)^{-1}, 1-\alpha_1, -\tfrac 12, -2, -1\big) \simeq M_3\big(2+(\alpha_3-1)^{-1}, 1-\alpha_1, \tfrac 12\big).\] We find that $2$ is exceptional on $M_5(\alpha)$ if and only if one of the following holds; $1-\alpha_1=3$, $(1-\alpha_1, 2+(\alpha_3-1)^{-1})=(4, \frac 12)$, $(\frac 52, \frac 32)$ . If $\alpha_2 = -\frac 12$ then \[M_5(\alpha)(\beta) \sim M_5\big(-2, -1, \tfrac 23, 1-\alpha_1, \alpha_3^{-1}\big) \simeq M_3\big(\tfrac 83, 1-\alpha_1, 1+\alpha_3^{-1}\big).\] We observe that $2$ is exceptional on $M_5(\alpha)$ if and only if one of the following holds; $1-\alpha_1=3$, $(1-\alpha_1, 1+\alpha_3^{-1}) = (4, \frac 12)$, $(\frac 32, \frac 52)$. If $\alpha_3 = \frac 23$ then \[M_5(\alpha)(\beta) \sim M_5\big(-2, 1-\alpha_2, -\tfrac 12, \alpha_1^{-1}, -1\big) \simeq M_3\big(2-\alpha_2, -\tfrac 12,2+\alpha_1^{-1} \big).\] We get that $2$ is exceptional on $M_5(\alpha)$ if and only if one of the following holds; $2+\alpha_1^{-1}=0$, $(1-\alpha_2, \alpha_1^{-1})=(-1, -1)$, $(2-\alpha_2, 2+\alpha_1^{-1})=(4, \frac 12)$, $(\frac 52, \frac 32)$. If $\alpha_3 = \frac 13$ then \[M_5(\alpha)(\beta) \sim M_5\big(-1, \alpha_2, \tfrac 13,1-\alpha_1^{-1}, -2\big) \simeq M_3\big(\alpha_2+2, \tfrac 13,2-\alpha_1^{-1}\big).\] The result is that $2$ is an exceptional slope on $M_5(\alpha)$ if and only if one of the following holds; $\alpha_2+2=0$, $(\alpha_2+2,2-\alpha^{-1})=(\frac 12, 4)$, $(\frac 32, \frac 52)$, $(-1, -1)$, $(-1, 4)$. Thus, the set of all $(-2, \alpha_1,\alpha_2,\alpha_3, 2)$ constructed in the above analysis is the set $l_2$. This completes the construction of $l\cup l_{-1}\cup l_{\frac 12}\cup l_2$ and the sets $\{(M_5(\alpha), E_{\tau}(M_5(\alpha))\}$. The last step is to reduce the size of $\{(M_5(\alpha), E_{\tau}(M_5(\alpha))\}$ using maps (\ref{first:eqn})-(\ref{figure8:eqn}). The result is shown in Tables \ref{tb1}-\ref{tb4}. The corresponding exceptional manifolds shown in Tables \ref{tb1}-\ref{tb4} are written down using the description of isolated exceptional fillings of $M_3$ and $M_5$ provided in \cite{Magic} and \cite{Big} respectively. $\square$ \vspace{.5cm} The proof of Theorem \ref{4CL_thm} follows a similar line of reasoning: \vspace{.4cm} \textsc{Proof of Theorem \ref{4CL_thm}. } We let $\gamma=(\gamma_0,\gamma_1, \gamma_2, \gamma_3)$ be a hyperbolic filling instruction on $M_4$ with at least on $\varnothing$ slope. We translate the problem on 4CL into a problem on 5CL; using the Rolfsen twist we see that $M_4(\gamma) \simeq M_5(\delta)$ where $\delta=(\gamma_0-1,\gamma_1, \gamma_2, \gamma_3-1,-1)$ (here $\gamma_0-1$, $\gamma_3-1$ should be read as $\varnothing$ when $\gamma_0$, $\gamma_3=\varnothing$). Now the argument proceeds exactly as in the proof of Theorem \ref{5CL_thm}. We use Identities (\ref{first:eqn})-(\ref{3rd:eqn}) to assume without loss of generality that $\delta_0=-1$ and that $\tau$ is the boundary corresponding to the neighbourhood of the 4th chain component of 5CL; namely we can assume $\delta=(\delta_0, \delta_1, \delta_2, \delta_3, \varnothing)$ is a hyperbolic filling instruction on $M_5$ that does not factor through $M_3$, and that some slope of $\delta$ is in $\{-1, \frac 12, 2\}$. The conditions of $\delta$ being non-exceptional and not factoring through $M_3$ means that $\delta_1\not\in\{0, 1, \infty,-1, -2, -\frac 12\}$ and $\delta_2, \delta_3\not\in\{0, 1, \infty, -1, \frac 32, 3\}$. Theorem \ref{main:teo} implies that every $\beta\in\{-1, 0, 1, \infty\}$ is an exceptional slope on $M_5(\delta)$, and so $\{-1, 0, 1, \infty\}\subseteq E_{\tau}(M_5(\delta))$ and $e_{\tau}(M_5(\delta))\geq 4$. Identities (\ref{4CLinf})-(\ref{4CL1}) follow directly from Identities (\ref{5CLinf})-(\ref{5CL0}) while (\ref{4CL0}) is established using (\ref{blow:eqn}) and then a composition of (\ref{first:eqn})-(\ref{3rd:eqn}). More generally, Theorem \ref{main:teo} implies that if $\beta$ is an exceptional slope on $M_5(\delta)$ then $\beta\in\{-1, 0, 1, \infty\}$, or $(\delta, \beta)$ factors through $M_3$, or $M_5(\delta, \beta)=m_2$. The $(\delta, \beta)$ coming from $m_2$ are given by $(-1, -3, -2, -2, -3)$, $(-1, \frac 13, \frac 43, \frac 43, \frac 13)$, $(-1, -\frac 13, 4, \frac 23, 3)$, $(-1, 3, \frac 23, 4, -\frac 13)$. This implies that $\beta$ is in one of $[\![(1)]\!]$, $[\![(-1)]\!]$, $[\![(-2)]\!]$, $[\![(-3)]\!]$. Every slope in $\{[\![(-1)]\!], [\![(-2)]\!], [\![(-3)]\!]\}$ is examined individually as in the proof of Theorem \ref{5CL_thm} to obtain a complete list of all $\big(M_5(\delta), E_{\tau}(M_5(\delta))\big)$ pairs that have $\delta=\big(-1, \frac pq, \frac rs, \frac uv, \varnothing\big)$ hyperbolic not factoring through $M_3$ with $e\big(M_5(\delta)\big)>4$. Lastly, the list of $\delta$ produced is reduced using the maps (\ref{first:eqn})-(\ref{figure8:eqn}), and a positive twist about the zeroth component of the 5CL gives the list of $\alpha$'s on $M_4$ and the exceptional slopes $\beta_i$ on $M_4(\alpha)$ shown in Table \ref{tb5}-\ref{tb6}. The corresponding exceptional manifolds shown in Tables \ref{tb5}-\ref{tb6} are written down using the description of exceptional fillings on $M_3$ and $M_5$ provided in \cite{Magic} and \cite{Big}. $\square$ \vspace{.5cm} We finish by pointing out that Theorems \ref{5CL_thm} and \ref{4CL_thm} imply that any one-cusped hyperbolic manifold $M$ obtained by surgery on 5CL with 10 exceptional slopes, or having two exceptional slopes at distance 8, is obtained by surgery on 3CL and that the only such $M$'s obtained by surgery on 3CL are the figure-8 knot exterior and its sister (see the Appendix in \cite{Magic}) which shows the final two assertions of Theorem \ref{flash}. The second assertion in Theorem \ref{flash} is a simple consequence of the fact that, up to (\ref{first:eqn})-(\ref{figure8:eqn}), the only $M_5(\alpha)$ with a pair of exceptional slopes at distance 4 with $\alpha$ not factoring through $M_3$ is $M_4(-2,-\tfrac12,-2)$ with the slopes $-1, 3$ (see Tables \ref{tb2}-\ref{tb1}), and $M_4(-2,-\tfrac12,-2)(-1)$, $M_4(-2,-\tfrac12,-2)(3)$ are not lens spaces; all $M_3(\alpha)$ with a pair of exceptional slopes at distance 4 can be found in \cite{Magic}, and it is not hard to check that the corresponding pair of exceptional fillings are never a lens space, toroidal pair. Lastly, we remark that the description of exceptional fillings provided in Theorems \ref{5CL_thm} and \ref{4CL_thm}, and \cite{Magic} can be used to easily enumerate all examples of exceptional pairs at maximal distance; one such enumeration shows that no counterexample to the Cabling conjecture is obtained by surgery on 5CL (see \cite{thesis} for full details). \newpage \section{Tables}\label{tables} \begin{table}[h!] \begin{center} \begin{tabular}{|c|c|c|} \hline & Additional & Exceptional\\ $\alpha$ & exceptional & filling $M_5(\alpha)(\beta_i)$\\ & slopes $\beta_i$ & \\ \hline \hline \phantom{\Big|}$(-2, -\frac{1}{2}, 3, 3)$ & $\beta_1=-1$ & $\seiftre{S^2}215{-2}4{-1}$\\ & $\beta_2=-\frac12$ &$\seifuno{A}2{-1}\bigb1211$ \\ \hline \phantom{\Big|}$(-2, \frac 32, \frac 32, -2)$ & $\beta_1=-1$ & $\seifdue{\mathbb{RP}^2}2{-1}31$\\ & $\beta_2=-\frac12$ & $\seifuno{A}2{-1}\bigb1211$ \\ \hline \phantom{\Big|}$(-2, -3, -\frac 12, -2)$& $\beta_1=-1$ & $\seiftre{S^2}2{-1}31{13}2$ \\ & $\beta_2=2$ & $\seifdue D2131\bigu{1}110 \seifdue D2{1}3{-8}$\\ \hline \phantom{\Big|}$(-2, -\frac{1}{3}, 3, \frac{2}{3})$ & $\beta_1=-1$ & $\seiftre{S^2}217{-2}5{-3}$\\ & $\beta_2=2$ & $\seifuno{A}2{3}\bigb0110$\\ \hline \phantom{\Big|}$(-2, -\frac{1}{2}, 3, \frac{2}{3})$ & $\beta_1=-1$ & $\seiftre{S^2}215{-2}5{-3}$\\ & $\beta_2=2$ & $\seiftre{S^2}2{-1}31{11}2$ \\ \hline \phantom{\Big|}$(-2, -2, -2, -2)$ & $\beta_1=-1$ & $\seiftre{S^2}2{-1}3171$\\ & $\beta_2=-2$ & $\seifdue D212{-1}\bigu{-1}51{-4} \seifdue D2{1}31$\\ \hline \end{tabular} \end{center} \caption{$\alpha$ a filling instruction on $M_5$ not factoring through $M_4$ with $e_{\tau}\big(M_5(\alpha)\big) = 5$, and $E_{\tau}\big(M_5(\alpha)\big) = \{\beta_1, \beta_2, 0,1,\infty\}$. \label{tb1}} \end{table} \begin{table}[h!] \begin{center} \begin{tabular}{|c|c|c|} \hline & Additional & Exceptional\\ $\alpha$ & exceptional & filling $M_5(\alpha)(\beta)$\\ & slopes $\beta$ & \\ \hline \hline \phantom{\Big|}$(-2, \frac{p}{q}, 3, \frac{u}{v})$ & $-1$ & $\seifdue D21p{-q}\bigu0110 \seifdue D2{1}{u+v}{-v}$\\ \hline $(-2, \frac{p}{q}, \frac{r}{s}, -2)$ & $-1$ & $\seifdue D{s}{2s-r}{q}{q-p}\bigu01{-1}{-1} \seifdue D2{1}{3}{1}$ \\ \hline \phantom{\Big|}$(-2, \frac{3}{2}, \frac{3}{2}, \frac{u}{v})$ & $-1$ &$\seifdue D2131\bigu110{-1} \seifdue D2{1}{u}{-v}$ \\ \hline \phantom{\Big|}$(-2, \frac{p}{q}, \frac{5}{2}, -\frac{1}{2})$ & $-1$ &$\seifdue D2131\bigu110{-1} \seifdue D21{q-p}{q}$ \\ \hline \phantom{\Big|}$(-2, -2, \frac{r}{s}, -3)$ & $-1$ &$\seifuno{A}s{s-r}\bigb0110$ \\ \hline \phantom{\Big|}$(-2, -\frac{1}{2}, 4, \frac{u}{v})$ & $-1$ &$\seifdue D2131\bigu1110 \seifdue D21{v}{-u-2v}$ \\ \hline \phantom{\Big|}$(-2, \frac{p}{q}, 4, -\frac{3}{2})$ & $-1$ &$\seifdue D2131\bigu1110 \seifdue D21{q}{-p-q}$\\ \hline \phantom{\Big|}$(-2, \frac{1}{3}, \frac{r}{s}, \frac{3}{2})$ & $\frac{1}{2}$ & $\seiftre{S^2}2{-1}31{5r-4s}{r-s}$ \\ \hline \end{tabular} \end{center} \caption{$\alpha$ a filling instruction on $M_5$ not factoring through $M_4$ with $e_{\tau}\big(M_5(\alpha)\big) =4$, and $E_{\tau}\big(M_5(\alpha)\big) = \{\beta, 0,1,\infty\}$, part 1/3. \label{tb2}} \end{table} \begin{table}[h!] \begin{center} \begin{tabular}{|c|c|c|} \hline & Additional & Exceptional\\ $\alpha$ & exceptional & filling $M_5(\alpha)(\beta)$\\ & slopes $\beta$ & \\ \hline \hline \phantom{\Big|}$(-2, 4, 5, -\frac{3}{2})$& $-1$ &$\seifuno{A}21\bigb0110$\\ $(-2, -\frac{1}{2}, 5, 3)$& & \\ \hline $(-2, 3, 4, -\frac{4}{3})$&$-1$&$\seifuno{A}21\bigb1110$\\ \hline $(-2, 3, \frac{3}{2}, -\frac{1}{2})$&$-1$ &$T\bigb{-3}1{-1}0$\\ \hline $(-2, -2, 4, -\frac{5}{3})$&$-1$&$\seifdue D2121\bigu01{-1}{-1} \seifdue D2131$\\ $(-2, -\frac{2}{3}, 4, -3)$& &\phantom{\Big|}\\ \hline $(-2, \frac{2}{3}, \frac{5}{2}, -\frac{1}{3})$&$-1$& $\seifdue D2121\bigu{-1}1{0}{-1} \seifdue D2131$ \phantom{\Big|}\\ \hline $(-2, \frac{4}{3}, \frac{3}{2}, \frac{1}{3})$&$-1$&$\seifuno{A}21\bigb1110$ \phantom{\Big|}\\ \hline $(-2, -3, -2, -3)$& $-1$ & $\seiftre{S^2}2{-1}3171$ \phantom{\Big|} \\ $(-2, -2, -2, -4)$& & \\ \hline \end{tabular} \end{center} \caption{$\alpha$ a filling instruction on $M_5$ not factoring through $M_4$ with $e\big(M_5(\alpha)\big) =4$, and $E_{\tau}\big(M_5(\alpha)\big) = \{\beta, 0,1,\infty\}$, part 2/3. \label{tb3}} \end{table} \begin{table}[h!] \begin{center} \begin{tabular}{|c|c|c|} \hline & Additional & Exceptional\\ $\alpha$ & exceptional & filling \\ & slope $\beta$ & $M_5(\alpha)(\beta)$ \\ \hline \hline $(-2, -4, -2, -3)$ & & \\ $(-2, -2, -2, -5)$ & $-1$ & $\seiftre{S^2}2{-1}4151$ \phantom{\Big|} \\ $(-2, -3, -3, -3)$ & & \\ $(-2, -2, -3, -4)$ & \phantom{\Big|} & \\ \hline $(-2, -2, -4, -4)$ & & \\ $(-2, -5, -2, -3)$& $-1$ & $\seiftre{S^2}3{-2}3151$ \phantom{\Big|} \\ $(-2, -3, -4, -3)$& & \\ $(-2, -2, -2, -6)$ & \phantom{\Big|} & \\ \hline $(-2, -2, -2, -7)$ & & \\ $(-2, -6, -2, -3)$ &$-1$&$\seifdue D2121\bigu0110 \seifdue D2131$ \phantom{\Big|} \\ $(-2, -3, -5, -3)$ & & \\ $(-2, -2, -5, -4)$ & \phantom{\Big|} & \\ \hline $(-2, -4, -3, -3)$ & $-1$ & $\seifdue D2121\bigu120{-1} \seifdue D2131$ \phantom{\Big|} \\ $(-2, -2, -3, -5)$& & \\ \hline $(-2, -3, -2, -4)$ & $-1$ & $\seifdue D2121\bigu23{-1}{-2} \seifdue D2131$ \phantom{\Big|}\\ \hline $(-2, \frac 32, \frac 52, -\frac 23)$ & $-1$ & $\seifuno{A}21\bigb2110$ \phantom{\Big|}\\ \hline $(-2, -2, \frac{1}{4}, 3)$& $\frac12$ & $\seiftre{S^2}213{-1}9{-2}$ \phantom{\Big|}\\ \hline $(-2, \frac 14, \frac 32, \frac 43)$ & $\frac12$ & $\seifdue D2121\bigu{-1}41{-3} \seifdue D2131$ \phantom{\Big|}\\ \hline $(-2, \frac 13, \frac 23, \frac 53) $ & $\frac12$ & $\seiftre{S^2}213{-1}52$ \phantom{\Big|} \\ \hline \end{tabular} \end{center} \caption{$\alpha$ a filling instruction on $M_5$ not factoring through $M_4$ with $e\big(M_5(\alpha)\big) =4$, and $E_{\tau}\big(M_5(\alpha)\big) = \{\beta, 0,1,\infty\}$, part 3/3. \label{tb4}} \end{table} \begin{table}[h!] \begin{center} \begin{tabular}{|c|c|c|} \hline & Additional & Exceptional\\ $\alpha$ & exceptional & filling \\ & slopes $\beta_i$ & $M_4(\alpha)(\beta_i)$ \\ \hline \hline $(-2, -2, -2)$ & $\beta_1=-2$ & $\seifdue D212{-1}\bigu{-1}41{-3} \seifdue D2{1}31$ \phantom{\Big|}\\ & $\beta_2=-1$ & $T_{{\tiny{\matr31{-1}0}}\phantom{\Big|}}\phantom{\Big|}$ \\ \hline $(-2, -\frac{1}{2}, -2)$ & $\beta_1=-1$ & $\seifuno{A}21\bigb{-1}110$ \phantom{\Big|} \\ & $\beta_2=3$ & $\seifdue D2121\bigu{-1}10{-1} \seifdue D2{1}31$ \\ \hline \end{tabular} \end{center} \caption{$\alpha$ a filling instruction on $M_4$ not factoring through $M_3$ with $e_{\tau}(M_4(\alpha)\big) = 6$, and $E_{\tau}\big(M_4(\alpha)\big) = \{\beta_1,\beta_2, 0,1,2,\infty\}$. \label{tb5}} \end{table} \begin{table}[h!] \begin{center} \begin{tabular}{|c|c|c|} \hline & Additional & Exceptional\\ $\alpha$ & exceptional & filling \\ & slope $\beta$ & $M_4(\alpha)(\beta)$ \\ \hline \hline $(-2, \frac{r}{s}, -2)$ & & $\seifuno{A}s{s-r}\bigb0110$ \phantom{\Big|}\\ $(\frac{p}{q}, 4, -\frac{1}{2})$ & & $\seifdue D2131\bigu1110 \seifdue D21{-q}{p+q}$ \phantom{\Big|}\\ $(4, 5, -\frac{1}{2})$ & & $\seifuno{A}21\bigb0110$ \phantom{\Big|}\\ $(-2, 4, -\frac{2}{3})$& & $\seifdue D212{-1}\bigu{-1}10{-1} \seifdue D2{1}31$ \phantom{\Big|}\\ $(-2, -5, -3)$& $\beta=-1$ &$\seifdue D212{1}\bigu0110 \seifdue D2{1}31$ \phantom{\Big|}\\ $(-2, -2, -6)$& &$\seifdue D212{1}\bigu0110 \seifdue D2{1}31$ \phantom{\Big|}\\ $(-2, -2, -3)$ & \phantom{\Big|} & $\seiftre{S^2}2{-1}3171$ \\ $(-2, -3, -3)$,& \phantom{\Big|} & $\seiftre{S^2}2{-1}4151$\\ $(-2,-2,-4)$ & \phantom{\Big|} & $\seiftre{S^2}2{-1}4151$ \\ $(-3, -4, -2)$&\phantom{\Big|} & $\seiftre{S^2}3{-2}3141$ \\ $(-2,-2,-5)$ &\phantom{\Big|} & $\seiftre{S^2}3{-2}3141$\\ $(-3, -2, -3)$ &\phantom{\Big|} & $\seifdue D212{-1}\bigu{-1}31{-2} \seifdue D2{1}31$\\ \hline \end{tabular} \end{center} \caption{$\alpha$ a filling instruction on $M_4$ not factoring through $M_3$ with $e_{\tau}(M_4(\alpha)\big) =5$, and $E_{\tau}\big(M_4(\alpha)\big) = \{\beta, 0,1,2, \infty\}$. \label{tb6}} \end{table} \clearpage
\section{\label{sec:level1}First-level heading} The anomalous behavior of the density of water as it transitions to ice and its associated hydrogen bond (defined as the entire O:H-O) angle-length-stiffness cooling relaxation dynamics continue to baffle the field, despite the intensive investigations carried out in the past decades \cite{supp,clark10a,soper10,head06,clark10b,petkov12,stone07, stokely10, huang09,english11,malla07a,malla7b,mishima98,moore11,molinero,wang09}. Established database \cite{malla07a} shows that both the liquid and the solid H$_2$O undergoes the seemingly regular process of cooling densification at different rates. At the water-ice transition phase, volume expansion takes place and results in ice with a density 9$\%$ lower than the maximal density of water at 277 K \cite{malla07a,supp}. The cooling densification is associated with a redshift of the high-frequency H-O phonons $\omega_H$ ($\sim$3000 cm$^{-1}$)\cite{cross37,yoshmura} while the freezing expansion is accompanied with a blueshift of the $\omega_H$ \cite{duric11}. However, the understanding of the fundamental nature underlying the observed mass-density and phonon-frequency transition dynamics and their correlation still remains incomplete. Numerous models have been developed for explaining water's expansion upon freezing or at supercooling state. The elegant models include the thermal fluctuations in the mono-phase of tetrahedrally-coordinated structures \cite{head06,petkov12} and the mixed-phase of low- and high- density fragments with thermal modulation of the fragmental ratios \cite{huang09,wernet04}. Little attention has been paid to the mechanism for the seemingly regular process of cooling densification in both the liquid and the solid phase. The Raman shift of the low-frequency O:H ($\omega_L \sim$ 200 cm$^{-1}$) phonons in various phases has not yet been systematically characterized. Therefore, there is a great need for a consistent framework to understand the density and phonon-frequency transitions and the associated hydrogen bond angle-length-stiffness relaxation dynamics of H$_2$O through its full set of states. \\ \indent In this letter, we show that we have been able to reconcile the measured mass-density \cite{malla07a,supp} and Raman-frequency transitions of water/ice based on the framework of O:H-O bond specific-heat disparity, Raman sectroscopy measurements, and molecular dynamics (MD) calculations of the hydrogen bond angle-length-stiffness relaxationin of water/ice over the full temperature range. The often overlooked Coulomb repulsion between the electron pair in the H-O covalent bond and the nonbonding electron lone pair of oxygen \cite{sun12} and the specific-heat disparity of the O:H-O bond are shown to be the key to resolving the density puzzles. \\ \indent We consider the basic structural unit of O$^{\delta -}$:H$^{\delta +}$-O$^{\delta -}$ \cite{sun12,sunprl} (also denoted as ``O$\cdots$H-O") to represent the O$^{\delta -}$-O$^{\delta -}$ interactions in H$_2$O, except for H$_2$O under extremely high pressures and temperatures \cite{wang11}. The fraction $\delta$ represents the polarity of the H$^{\delta +}$-O$^{\delta -}$ polar-covalent bond. In Fig.\ref{fig1a}, the pair of dots on the left represents the nonbonding lone pair ``:" and the pair on the O atom on the right represents the bonding electron pair ``-". The H atom serves as the point of reference in the O:H-O system. The lone pair on the left belongs to the sp$^3$-orbit hybridized oxygen and the bonding pair is shared by the H-O and centred at sites close to oxygen. For completeness, we define the entire hydrogen bond to be O:H-O, the intra-molecular polar-covalent bond as the H-O bond and the inter-molecular van der Waals (vdW) bond as the O:H bond hereon. \floatsetup[figure]{style=plain,subcapbesideposition=top} \begin{figure}[!hbtp] \sidesubfloat[]{\includegraphics[width=3.2in, trim=150 250 360 110, clip]{fig_1a} \label{fig1a}}\\ \sidesubfloat[]{\includegraphics[width=3.0in, trim=0 0 0 30, clip]{fig_1b} \label{fig1b}} \vspace{-0.8em} \caption{(a, b) Forces and (c) the respective specific heats of the two segments in the O:H-O bond. Combined with the Coulomb repulsion $f_q$ between the electron pairs (pairs of dots in a) and the resistance to deformation $f_{rx}$ ($x = H$ for the H-O and $L$ for the O:H segment), the cooling contraction force $f_{dx}$ drives the two segments to relax in the same direction but by different amounts. (c) Because of the difference in their Debye temperatures (Table \ref{tab1}), the specific heat $\eta_L$ of O:H rises faster towards the asymptotic maximum value than the $\eta_H$. Such a specific-heat decomposition results in three regions that correspond, respectively, to the phases of liquid (I), solid (III), and liquid-solid transition (II) with different specific-heat ratios. R is the gas constant. (The $\eta_L$ in the solid phase differs from the $\eta_L$ in the liquid, which does not influence the validity of the hypothesis).} \label{fig1} \end{figure} As illustrated in Fig.\ref{fig1a}, a hydrogen bond is comprised of the O:H bond (broken lines) and the H-O bond rather than either of them alone. The H-O bond is much shorter, stronger, and stiffer than the O:H bond (comparison shown in Table \ref{tab1}). The O:H bond breaks at the evaporating point (T$_v$) of water (373 K)\cite{cross37}. However, the H-O bond is much harder to break as the bond energy of $\approx$4.0 eV \cite{supp} is twice that of the C-C bond in diamond (1.84 eV) \cite{kittel}. \begin{table*}[!htbp] \caption{Summary of the segmental length d$_x$, strength E$_x$ (energy)\cite{supp}, Debye temperature $\Theta_{Dx}$ \cite{zhao07,suter06}, stiffness $\omega_{x}$ (vibration frequency), melting point T$_{mx}$, and the inter-atomic and inter-electron-pair interactions of the O:H-O bond compared with those of the C-C bond in a diamond \cite{kittel}. } \centering \begin{tabular}{ccccccc} \hline\hline Segment (x) & d$_x$ (nm) & E$_x$ (eV) & $\omega_x$ (cm$^{-1}$) & $\Theta_{Dx}$ (K) & T$_{mx}$ (K) & Interaction \\ \hline H-O & $\sim$0.10 & $\sim$4.00 & $>$3000 & $>$3000 & - & Exchange \\ O:H & $\sim$0.20 & $\sim$0.05-0.10 & $\sim$200 & 198 & 373 & vdW \\ O--O & - & - & - & - & - & Repulsion \\ C-C & 0.15 & 1.84 & 1331 & 2230 & 3800 & Exchange \\ \hline \end{tabular} \label{tab1} \end{table*} Combined with the forces of the Coulomb repulsion, $f_q$$\propto$(d$_{O-O}$)$^{-2}$, and resistance to deformation, $f_{rx}$($x = H$ for the H-O and $L$ for the O:H bond), the force of cooling contraction, $f_{dx}$ drives these two segments to relax in the same direction but by different amounts. The $k_x$, which varies nonlinearly with the $d_x$, approximates the second derivative of the inter-atomic potential at equilibrium. The magnitude of the $f_{dx}$ varies with the specific heat of the specific bond in a particular temperature range. The Coulomb repulsion between the electron pairs, as represented by the pairs of dots in Fig.\ref{fig1a}, is the key to the O:H-O bond relaxation under excitation \cite{supp,sun12}. \\ \indent Generally, the specific heat is regarded as a macroscopic quantity integrated over all bonds, and is the amount of energy required to raise the temperature of the substance by 1 K. However, in dealing with the representative bond of the entire specimen, one has to consider the specific heat per bond that is obtained by dividing the bulk specific heat by the total number of bonds involved. In the case of the O:H-O bond, we need to consider the specific heat ($\eta_x$) characteristics of the two segments separately (see Fig.\ref{fig1b}) because of the difference in their strengths. The slope of the specific-heat curve is determined by the Debye temperature ($\Theta_{Dx}$) while the integration of the specific-heat curve from 0 K to the T$_{mx}$ \cite{sun09} is determined by the cohesive energy of the bond energy E$_x$. The specific-heat curve of the segment with a relatively lower $\Theta_{Dx}$ value will rise to the maximum value faster than the other segment. The $\Theta_{Dx}$, which is lower than the T$_{mx}$, is proportional to the characteristic frequency of the vibration ($\omega_x$) of the segment. Thus, we have the following relations (see Table \ref{tab1}): \begin{equation} \label{eq1} \left\{ \begin{array}{l} \frac{\Theta_{DL}}{\Theta_{DH}} \approx \frac{198}{\Theta_{DH}} \approx \frac{\omega_L}{\omega_H} \approx \frac{200}{3000} \approx \frac{1}{15} \\ \frac{\left(\int^{T_{mH}}_0 \eta_H dt \right)}{\left(\int^{T_{mL}}_0 \eta_L dt \right)} \approx \frac{E_H}{E_L} \approx \frac{4.0}{0.1} \approx 40 \end{array} \right. \end{equation} Based on this consideration, the maximal specific-heat ratio is estimated to be $\frac{\eta_H}{\eta_L}$$\approx$8. Such a specific-heat disparity between the O:H and the H-O segments creates three temperature regions with different $\frac{\eta_L}{\eta_H}$ ratios, which should correspond to the phases of liquid (I), solid (III), and liquid-solid transition (II). \\ \indent The consistency in the number of regions (i.e. phases I, II, III) of the proposed specific-heat curve (Fig.\ref{fig1b}), the mass-density transition \cite{malla07a}, and the O:H-O bond angle-length-stiffness relaxation dynamics (Fig. 2 and Fig. 3) suggest that the segment with a relatively lower specific heat is thermally more active than the other segment. This thermally active segment serves as the ``master'' that undergoes cooling contraction while forcing the other ``slave'' segment to elongate through Coulomb repulsion. Therefore, as can be derived from Fig.\ref{fig1b}, the specific-heat ratio, the master segment, and the O--O length change in each temperature region are correlated as follows (see \cite{supp} for details): \vspace{-1.0em} \begin{widetext} \begin{equation} \label{eq2} \left. \begin{array}{l} \text{II} \\ \text{I, III} \\ \text{Transition} \end{array} \begin{array}{l} (\eta_H < \eta_L): \\ (\eta_L < \eta_H): \\ (\eta_H = \eta_L): \end{array} \begin{array}{l} f_{dH}>(f_{dL} + f_{rL} + f_{rH}) \\ f_{dL}>(f_{dH} + f_{rL} + f_{rH})\\ f_{dH}=f_{dL} \end{array} \right\} \Rightarrow \Delta d_{O-O} \Rightarrow \left(\Delta V \right)^{1/3} \left\{ \begin{array}{l} <\\ >\\ = \end{array} \right\} 0 \end{equation} \end{widetext} \noindent Because of the strength difference of the two segments \cite{sun12}, the length of the softer O:H bond always relaxes more than that of the stiffer H-O bond: $|\Delta d_L|$$>$$|\Delta d_H|$. The two segments relax in the same direction because of the repulsion. Thus, we expect the O:H-O bond to relax in the following manners during cooling:\\ i) In the transition phase II, $\eta_H$$<$$\eta_L$ and $f_{dH}$$\gg$$f_{dL}$. The H-O bond contraction dominates. The stiffer H-O bond contracts less than the O:H bond elongates, resulting in $\Delta d_{O-O}$=$\Delta d_L - \Delta d_H$$>$0. Therefore, a net O--O length gain and an accompanying volume expansion ($\Delta$V $>$ 0) takes place. \\ ii) In the liquid I and the solid III phase, $\eta_L$$<$$\eta_H$ and $f_{dL}$$\gg$$f_{dH}$. The master and the slave swap roles. The softer O:H bond contracts significantly more than the H-O bond elongates, $\Delta d_{O-O}$=$\Delta d_H - \Delta d_L$$<$0. Hence, a net O--O contraction results in a gain in the mass density. \\ iii) At the crossing points (Fig.\ref{fig1b}), $\eta_H$$=$$\eta_L$ and $f_{dH}$=$f_{dL}$. There is a transition between O--O expansion and contraction, corresponding to the density maximum at 277 K and the density minimum below the freezing point \cite{malla07a,malla7b}. \\ iv) Meanwhile, the repulsion increases the O:H-O angle $\theta$ and polarizes the electron pairs during relaxation. \\ \indent It has been shown that a segment increases in stiffness as it becomes shorter, while the opposite occurs as it elongates \cite{sun12,li12}. The Raman shift, which is proportional to the square root of bond stiffness, approximates the length and strength change of the bond during relaxation directly. Approximating the vibration energy of a vibration system to the Taylor series of the inter-atomic potential energy, u$_x$(r), leads to: \vspace{-0.3em} \begin{equation} \label{eq3} \Delta \omega_x \propto \left( \frac{\partial^2 u_x (r)}{\mu \partial r^2} \bigg|_{r=d_x} \right)^{1/2} \propto \frac{\sqrt{E_x/ \mu}}{d_x} \propto \sqrt{Y_x d_x} \end{equation} \noindent The stiffness is the product of the Young's modulus (Y$_x$$\propto$E$_x$/d$^3_x$) and the length of the segment in question \cite{sun12}. The $\mu$ is the reduced mass of the vibrating dimer. Therefore, the Raman shift is a measure of the segmental stiffness. \\ \indent In order to verify our hypotheses and predictions regarding the O:H-O bond angle-length-stiffness change, the specific-heat disparity, and the density and phonon-frequency anomalies of water/ice, we have conducted Raman measurements and MD calculations as a function of temperature. Two MD computational methods were used in examining the mono- and the mixed-phase models. Details of the experimental and calculation procedures are described in the supplementary information \cite{supp}. \floatsetup[figure]{style=plain,subcapbesideposition=top} \begin{figure}[!hbtp] \sidesubfloat[]{\includegraphics[width=1.55in, trim=0 50 40 20, clip]{fig_2a} \label{fig2a}} \sidesubfloat[]{\includegraphics[width=1.6in, trim=10 18 00 20, clip]{fig_2b} \label{fig2b}}\\ \sidesubfloat[]{\includegraphics[width=1.45in, trim=220 600 200 100, clip]{fig_2c} \label{fig2c}} \sidesubfloat[]{\includegraphics[width=1.5in, trim=0 10 115 50, clip]{fig_2d} \label{fig2d}}\\ \caption{(a) Segmental length change of the O:H-O bond in the phases of liquid (I), solid (III), and liquid-solid transition (II). Arrows denote the cooling contraction of the master segments, which are coupled with the expansion of the slaves. $\Delta T$ = T - T$_{max}$ with T$_{max}$ = 277 K is the maximal density temperature. (b) O:H-O bond angle widening driven by cooling also exhibits three regions. (c) Snapshots of the MD trajectory show that the V-shaped H-O-H molecues remain intact at 300 K because of the robustness of the H-O bond ($\sim$4.0 eV/bond) with pronounced quantum fluctuations in the angle and in the d$_L$ in liquid phase. (d) The change of the O--O distance agrees with the measured three-region water and ice densities \cite{supp,malla07a}. In ice, the O--O distance is longer than that in water, which results in ice floating.} \label{fig2} \end{figure} Fig.\ref{fig2} shows the MD-derived change of (a) the H-O bond and the O:H bond length, (b) the O:H-O bond angle $\theta$, (c) the snapshots of the MD trajectory, and (d) the O--O distance as a function of temperature. As shown in Fig.\ref{fig2a}, the shortening of the master segments (denoted with arrows) is always coupled with a lengthening of the slaves during cooling. The temperature range of interest consists of three regions: in the liquid region I and the solid region III, the O:H bond contracts significantly more than the H-O bond elongates, resulting in a net loss of the O--O length. Thus, cooling-driven densification of H$_2$O happens in both the liquid and the solid phase. This mechanism differs completely from the mechanism conventionally adopted for the standard cooling densification of other regular materials in which only one kind of chemical bond is involved. In other materials, cooling shortens and stiffens all the inter-atomic bonds \cite{sun05}. In contrast, in the transition phase II \cite{malla07a,mishima98,moore11}, the master and the slave swap roles. The O:H bond elongates more than the H-O bond shortens so that a net gain in the O--O length occurs. \\ \indent The $\theta$ angle widening (Fig.\ref{fig2b}) could also contribute to the volume change. In the liquid phase I, the mean $\theta$ valued at 160$^\circ$ remains almost constant. However, the snapshots of the MD trajectory in Fig.\ref{fig2c} and the MD movie in the attached \cite{MD movie} show that the V-shaped H-O-H molecules remain intact at 300 K over the entire duration recorded, accompanied by high fluctuations in the $\theta$ and the $d_L$ in this regime, which indicates the dominance of tetrahedrally-coordinated water molecules \cite{kuhne13}. In region II, cooling widens the $\theta$ from 160$^\circ$ to 167$^\circ$, which contributes a maximum of +1.75$\%$ to the O:H-O bond elongation and ~5.25$\%$ to volume expansion. In phase III, the $\theta$ increases from 167$^\circ$ to 174$^\circ$ and this trend results in a maximal value of -2.76$\%$ to the volume contraction. \\ \indent The calculated temperature dependence of the O--O distance as shown in Fig.\ref{fig2d} matches satisfactorily with that of the measured density profile \cite{supp,malla07a}. Importantly, the O--O distance is longer in ice than it is in water, and therefore, ice floats. \floatsetup[figure]{style=plain,subcapbesideposition=top} \begin{figure}[!hbtp] \sidesubfloat[]{\includegraphics[width=1.5in, trim=95 25 62 90, clip]{fig_3a} \label{fig3a}} \sidesubfloat[]{\includegraphics[width=1.46in, trim=105 25 60 100, clip]{fig_3b} \label{fig3b}} \vspace{-0.8em} \caption{Temperature dependent Raman shifts of (a) $\omega_L$ (the O:H phonon) and (b) $\omega_H$ show three regions: T$>$ 273 K (I), 273 K$<$T$<$258 K, and T$<$258 K (III) .} \label{fig3} \end{figure} \indent The measured Raman spectra in Fig.\ref{fig3} show three regions: T$>$273 K (I), 273 $\geq$ T $\geq$ 258 K (II), and T$<$258 K (III), which are in agreement with the MD calculations \cite{supp} and our predictions: \\ i) At T$>$273 K (I), abrupt shifts of the $\omega_L$ from 75 to 220 cm$^{-1}$ and the $\omega_H$ from 3200 to 3150 cm$^{-1}$ indicate ice formation. The coupled $\omega_L$ blueshift and $\omega_H$ redshift indicate that cooling shortens and stiffens the O:H bond but lengthens and softens the H-O bond in the liquid phase, which confirms the predicted master role of the O:H bond. \\ ii) At T$<$258 K (III), the trend of phonon relaxation remains the same as it is in the region of T $>$ 273 K despite a change in the relaxation rates. Cooling from 258 K stiffens the $\omega_L$ from 215 to 230 cm$^{-1}$ and softens the $\omega_H$ from 3170 to 3100 cm$^{-1}$. Other supplementary peaks at $\sim$300 and $\sim$3400 cm$^{-1}$ are found to be insignificant. The cooling softening of the $\omega_H$ mode agrees with that measured using IR spectroscopy \cite{medcrafy13} of ice clusters of 8$\sim$150 nm sizes. When the temperature drops from 209 to 30 K the $\omega_H$ shift from 3253 to 3218 cm$^{-1}$. For clusters of 5 nm size or smaller, the $\omega_H$ shifts by an addition of 40 cm$^{-1}$. \\ iii) At 273$\sim$258 K (II), the situation reverses. Cooling shifts the $\omega_H$ from 3150 to 3170 cm$^{-1}$ and the $\omega_L$ from 220 to 215 cm$^{-1}$, see the shaded areas. Agreeing with the Raman $\omega_H$ shift measured in the temperature range between 270 and 273 K \cite{supp,duric11}, the coupling of the $\omega_H$ blueshift and the $\omega_L$ redshift confirms the exchange in the master and the slave role of the O:H and the H-O bond during freezing. \\ \indent The MD-movie \cite{MD movie} shows that in the liquid phase, the H$^{\delta +}$ and the O$^{\delta -}$ attract each other between the H$^{\delta +}$:O$^{\delta-}$ but the O$^{\delta -}$-O$^{\delta -}$ repulsion prevents this occurrence. The intact O-H-O motifs are moving restlessly because of the high fluctuations and frequent switching of the H$^{\delta +}$:O$^{\delta -}$ interactions. Furthermore, the coupled cooling $\omega_L$ blueshift and $\omega_H$ redshift provide further evidence for the persistence of the Coulomb repulsion between the bonding and the nonbonding electron pairs in liquid. The presence of the electron lone pair results from the sp$^3$-orbit hybridization of oxygen that tends to form tetrahedral bonds with its neighbors \cite{supp,kuhne13}. Therefore, the H$¬_2$O in the bulk form of liquid could possesses the tetrahedrally-coordinated structures with thermal fluctuation \cite{petkov12,kuhne13,pauling35}. Snapshots of the MD trajectory in \cite{supp} revealed little discrepancy between the mono- and the mixed-phase structural models. \\ \indent The proposed mechanisms for: i) the seemingly regular processes of cooling densification of the liquid and the solid H$_2$O, ii) the abnormal freezing expansion, iii) the floating of ice, and, iv) the three-region O:H-O bond angle-length-stiffness relaxation dynamics of water and ice have been justified. Agreement between our calculations and the measured mass-density \cite{malla07a} and phonon-frequency relaxation dynamics in the temperature range of interest has verified our hypotheses and predictions:\\ i) Inter-electron-pair Coulomb repulsion and the segmental specific-heat disparity of the O:H-O bond determine the change in its angle, length and stiffness and the density and the phonon-frequency anomalies of water ice. \\ ii) The segment with a relatively lower specific-heat contracts and drives the O:H-O bond cooling relaxation. The softer O:H bond always relaxes more in length than the stiffer H-O bond does in the same direction. The cooling widening of the O:H-O angle contributes positively to the volume expansion at freezing.\\ iii) In the liquid and the solid phase, the O:H bond contracts more than the H-O bond elongates, resulting in the cooling densification of water and ice, which is completely different from the process experienced by other regular materials. \\ iv) In the freezing transition phase, the H-O bond contracts less than the O:H bond lengthens, resulting in expansion during freezing. \\ v) The O--O distance is larger in ice than it is in water, and therefore, ice floats. \\ vi) The segment increases in stiffness as it shortens, while the opposite occurs as it elongates. The density variation of water ice is correlated to the incoporative O:H and H-O phonon-frequency relaxaion dynamics. Special thanks to Phillip Ball, Yi Sun, Buddhudu Srinivasa, and John Colligon for their comments and expertise. Financial support received from NSF (Nos.: 21273191, 1033003, and 90922025) China is gratefully acknowledged.
\section{Introduction} A \textit{translation plane} is often represented by an algebraic structure called a \textit{quasifield.} Many properties of the translation plane can be most easily understood by looking at the appropriate quasifield. However, isomorphic translation planes can be represented by nonisomorphic quasifields. Furthermore, the collineations do not always have a nice representation in terms of operations in the quasifields. Let $p$ be a prime number and $(Q,+,\cdot)$ be a quasifield of finite order $p^n$. We identify $(Q,+)$ with the vector group $(\mathbb{F}_p^n, +)$. With respect to the multiplication, the set $Q^*$ of nonzero elements of $Q$ form a \textit{loop.} The \textit{right multiplication maps} of $Q$ are the maps $R_a:Q\to Q$, $xR_a=x\cdot a$, where $a,x\in Q$. By the right distributive law, $R_a$ is a linear map of $Q=\mathbb{F}_p^n$. Clearly, $R_0$ is the zero map. If $a\neq 0$ then $R_a\in \GL(n,p)$. In geometric context, the set of right translations are also called the \textit{slope set} or the \textit{spread set} of the quasifield $Q$, cf. \cite[Chapter 5]{Handbook}. The \textit{right multiplication group} $\rmlt(Q)$ of the quasifield $Q$ is the linear group generated by the nonzero right multiplication maps. It is immediate to see that $\rmlt(Q)$ is a transitive linear group, that is, it acts transitively on the set of nonzero vectors of $Q=\mathbb{F}_p^n$. The complete classification of finite transitive linear groups is known, the proof relies on the classification theorem of finite simple groups. Roughly speaking, there are four infinite classes and $27$ exceptional constructions of finite transitive linear groups. The first question this paper deals with asks which finite transitive linear group can occur as right multiplication group of a finite quasifield. It turns out that some of the exceptional finite transitive linear groups may happen to be the right multiplication group of a quasifield. Since these groups are relatively small, in the second part of the paper, we are able to give an explicit classification of all quasifields whose right multiplication group is an exceptional finite linear group. These results are obtained with computer calculations using the computer algebra system GAP4 \cite{GAP4} and the program CLIQUER \cite{Cliquer}. Right multiplication groups of quasifields have not been studied intensively. The most important paper in this field is \cite{Kallaher} by M. Kallaher, containing results about finite quasifields with solvable right multiplication groups. Another related paper is \cite{NagyG} which deals with the group generated by the left and right multiplication maps of finite semifields. Finally, we notice that our results can be interpreted in the language of the theory of finite loops, as well. \textit{Loops} arise naturally in geometry when coordinatizing point-line incidence structures. Most importantly, the multiplicative structure $(Q^*,\cdot)$ of $Q$ is a loop. In fact, any finite loop $\hat Q$ gives rise to a quasifield, provided the right multiplication maps of $\hat{Q}$ generate a group which is permutation isomorphic to a subgroup of $\GL(n,q)$, where the latter is considered as a permutation group on the nonzero vectors of $\mathbb{F}_q^n$. Therefore in this paper, we investigate loops whose right multiplication groups are contained is some finite linear group $\GL(n,q)$. \medskip There are many excellent surveys and monographs on translation planes and quasifields, see \cite{HughesPiper, Handbook, Luneburg} and the references therein. Our computational methods have similarities with those in \cite{CharnesDempwolff,Dempwolff}. \section{Translation planes, spreads and quasifields} \label{sec:geom} For the shake of completeness, in this section we repeat the definitions of concepts which are standard in the theory of translation planes. Moreover, we briefly explain the relations between the automorphisms of these mathematical objects. \medskip Let $\Pi$ be a finite projective plane. The line $\ell$ of $\Pi$ is a translation line if the translation group with respect to $\ell$ acts transitively (hence regularly) on the set of points $\Pi\setminus \ell$. Assume $\ell$ to be a translation line and let $\Pi^\ell$ be the affine plane obtained from $\Pi$ with $\ell$ as the line at infinity. Then $\Pi^\ell$ is a translation plane. By the theorems of Skornyakov-San Soucie and Artin-Zorn \cite[Theorem 6.18 and 6.20]{HughesPiper}, $\Pi$ is either Desarguesian or contains at most one translation line. This means that two finite translation planes are isomorphic if and only if the corresponding projective planes are. The relation between translation planes and quasifields is usually explained using the notion of \textit{(vector space) spreads.} \begin{definition} Let $V$ be a vector space over the field $F$. We say that the collection $\sigma$ of subspaces is a \emph{spread} if (1) $A,B \in \sigma$, $A\neq B$ then $V=A\oplus B$, and (2) every nonzero vector $x\in V$ lies in a unique member of $\sigma$. The members of $\sigma$ are the \emph{components} of the spread. \end{definition} If $\sigma$ is a spread in $V$ then Andr\'e's construction yields a translation plane $\Pi(\sigma)$ by setting $V$ as the set of points and the translates of the components of $\sigma$ as the set of lines of the affine plane. Conversely, if $\Pi$ is a finite translation plane with origin $O$ then we identify the point set of $\Pi$ with the group $\mathcal{T}(\Pi)$ of translations. As $\mathcal{T}(\Pi)$ is an elementary Abelian $p$-group, $\Pi$ becomes a vector space over (some extension of) $\mathbb{F}_p$ and the lines through $O$ are subspaces forming a spread $\sigma(\Pi)$. Andr\'e's construction implies a natural identification of the components of the spread with the parallel classes of the affine lines, and the points at infinity of the corresponding affine plane. The approach by spreads has many advantages. For us, the most important one is that they allow explicit computations in the group of collineation of the translation plane. Let $\Pi$ be a nondesarguesian translation plane and let us denote by $\mathcal{T}(\Pi)$ the group of translations of $\Pi$. The full group $\Aut(\Pi)$ of collineations contains $\mathcal{T}(\Pi)$ as a normal subgroup. Up to isomorphy, we can choose a unique point of origin $O$ in $\Pi$. The stabilizer $\mathcal{C}_O(\Pi)$ of $O$ in $\Aut(\Pi)$ is the \textit{translation complement} of $\Pi$ with respect to $O$. The full group $\Aut(\Pi)$ of collineations is the semidirect product of $\mathcal{T}(\Pi)$ and $\mathcal{C}_O(\Pi)$. In particular, $\mathcal{C}_O(\Pi)$ has the structure of a linear group of $V$. By \cite[Theorem 2.27]{Handbook}, the collineation group $\mathcal{C}_O(\Pi)$ is essentially the same as the group of automorphisms of the associated spread, where the latter is defined as follows. \begin{definition} Let $\sigma$ be a spread in the vector space $V$. The automorphism group $\Aut(\sigma)$ consists of the additive mappings of $V$ that permutes the components of $\sigma$ among themselves. \end{definition} As the translations act trivially on the infinite line, the permutation action of $\Aut(\sigma)$ is equivalent with the action of $\Aut(\Pi)$ on the line at infinity. Spreads are usually represented by a spread set of matrices. Fix the components $A,B$ of the spread $\sigma$ with underlying vector space $V$. The direct sum decomposition $V=A\oplus B$ defines the projections $p_A:V\to A$, $p_B:V\to B$. As for any component $C\in \sigma\setminus\{A,B\}$ we have $A\cap C=B\cap C=0$, the restrictions of $p_A$ and $p_B$ to $C$ are bijections $C\to A$, $C\to B$. Therefore, the map $u_C:A\to B$, $xu_C=(xp_A^{-1})p_B$ is a linear bijection from $A$ to $B$. When identifying $A,B$ with $\mathbb{F}_p^k$, $u_C$ can be given in matrix form $U_C$. The set $\mathcal{S}(\sigma)=\{U_C \mid C \in \sigma\}$ is called the \textit{spread set of matrices representing $\sigma$ relative to axes $(A,B)$.} This representation depends on the choice of $A,B\in \sigma$. It is also possible to think at a spread set as the collection $\mathcal{S}'(\sigma)=\{u_Du_C^{-1} \mid D \in \sigma \}$ of linear maps $A\to A$, with fixed $C$. A spread set $\mathcal{S}\subseteq \hom(A,B)$ can be characterized by the following property: For any elements $x\in A \setminus \{0\}$, $y\in B\setminus \{0\}$, there is a unique map $u\in \mathcal{S}$ such that $xu=y$. Indeed, if $\mathcal{S}=\mathcal{S}(\sigma)$ then $u=u_C$ where $C$ is the unique component containing $x\oplus y$. Conversely, $\mathcal{S}$ defines the spread \[ \sigma(\mathcal{S})=\{0\oplus B, A \oplus 0\} \cup \{ \{ x\oplus xu \mid x \in A \} \mid u \in \mathcal{S} \} \] with underlying vector space $V=A \oplus B$. \begin{definition} The autotopism group of the spread set $\mathcal{S}$ of $k\times k$ matrices over $F$ consists of the pairs $(T,U)\in \GL(k,F) \times \GL(k,F)$ such that $T^{-1}\mathcal{S}U=\mathcal{S}$. \end{definition} \cite[Theorems 5.10]{Handbook} says that autotopisms of spread sets relative to axes $(A,B)$ and automorphism of spreads fixing the components $A,B$ are essentially the same. \medskip By fixing a nondegenerate quadrangle $o,e,x,y$, any projective plane can be coordinatized by a \textit{planar ternary ring} (PTR), see \cite{HughesPiper}. Let $\Pi$ be a translation plane and fix affine points $o,e$ and infinite points $x,y$. Then, the coordinate PTR becomes a \textit{quasifield.} \begin{definition} The finite set $Q$ endowed with two binary operations $+$, $\cdot$ is called a \emph{finite (right) quasifield,} if \begin{enumerate} \item[(Q1)] $(Q,+)$ is an Abelian group with neutral element $0\in Q$, \item[(Q2)] $(Q\setminus\{0\},\cdot)$ is a loop, \item[(Q3)] the right distributive law $(x+y)z=xz+yz$ holds, and, \item[(Q4)] $x\cdot 0=0$ for each $x\in Q$. \end{enumerate} \end{definition} The link between tranlation planes and quasifields can be extended to spread sets, as well. In fact, the set $\mathcal{S}(Q)=\{R_x \mid x\in Q\}$ of nonzero right multiplication maps of $Q$ is a spread set relative to the infinite points of the $x$- and $y$-axes of the coordinate system. Collineations correspond to \textit{autotopisms} of the quasifield. \begin{definition} Let $(Q,+,\cdot)$ be a quasifield, $S,T,U:Q\to Q$ bijections such that $S,U$ are additive and $0T=0$. The triple $(S,T,U)$ is said to be an \emph{autotopism} of $Q$ if for all $x,y\in Q$, the identity $xS\cdot yT=(x\cdot y)U$ holds. \end{definition} It is easy to see that the triple $(S,T,U)$ is an autotopism of the quasifield $Q$ if and only if the pair $(S,U)$ is an autotopism of the associated spread set $\mathcal{S}(Q)$. We summarize the above considerations in the next proposition. \begin{proposition} \label{pr:collgrps} Let $\Pi$ be a translation plane, $\sigma$ the associated spread. Let $A,B$ be fixed components of $\sigma$ and $a,b$ the associated infinite points of $\Pi$. Let $\mathcal{S}$ be the spread set of $\sigma$ relative to axes $(A,B)$. Let $c,d$ be arbitrary affine points of $\Pi$ such that $a,b,c,d$ are in general position, and let $(Q,+,\cdot)$ be the coordinate quasifield of $\Pi$ with respect to the quadrilateral $abcd$. Then the following groups are isomorphic. \begin{enumerate} \item The autotopism group of $\mathcal{S}$. \item The stabilizer subgroup of the components $A,B$ in $\Aut(\sigma)$. \item The stabilizer subgroup of the triangle $abc$ in the full collineation group $\Aut(\Pi)$. \item The autotopism group of $Q$. \end{enumerate} \end{proposition} In particular, the structure of the autotopism group of the coordinate quasifield does not depend on the choice of the base points $c,d$. \section{Isotopy, parastrophy and computation} When investigating the isomorphism between translation planes, the \textit{isotopy of quasifields} is a central concept. We borrow the concept of \textit{parastrophy} from the theory of loops in order to define a wider class of equivalence for quasifields. \begin{definition} Let $\mathcal{S}, \mathcal{S}'$ be spread sets of matrices in $\GL(d,p)$. We say that $\mathcal{S}, \mathcal{S}'$ are \begin{enumerate} \item \emph{isotopes} if there are matrices $T,U \in\GL(d,p)$ such that $T^{-1}\mathcal{S}U=\mathcal{S}'$ holds. \item \emph{parastrophes} if there are matrices $T,U \in\GL(d,p)$ such that $T^{-1}\mathcal{S}U=\mathcal{S}'$ or $T^{-1}\mathcal{S}U=(\mathcal{S}')^{-1}$ holds. \end{enumerate} Analogously, we say that the quasifields $Q,Q'$ are isotopic (parastrophic) if their sets of nonzero right multiplications of matrices are isotopic (parastrophic) as spread sets of matrices. \end{definition} In Section \ref{sec:geom}, we explained the method of obtaining the spread set $\mathcal{S}$ of matrices from the spread $\sigma$ by fixing the components $A,B$. It follows that interchanging $A$ and $B$, the resulting spread set of matrices will be $\mathcal{S}^{-1}=\{u^{-1} \mid u \in \mathcal{S} \}$. Hence, $\mathcal{S}$ and $\mathcal{S}^{-1}$ determine isomorphic translation planes. Taking into account \cite[Propositions 5.36 and 5.37]{Handbook}, we obtain that parastrophic quasifields (or spread sets) determine isomorphic translation planes. On the one hand, the next proposition shows that the right multiplication group of a quasifield is parastrophy invariant. On the other hand, it will give us an effective method for the computation of parastrophy for quasifields with ``small'' right multiplication group. \begin{lemma} \label{lm:normal} Let $\mathcal{S}_1, \mathcal{S}_2$ be spread sets of matrices in $\GL(d,p)$. Assume that $1 \in \mathcal{S}_1, \mathcal{S}_2$ and define the transitive linear groups $G_1=\langle \mathcal{S}_1 \rangle$ and $G_2=\langle \mathcal{S}_2 \rangle$ generated by the spread sets. If $(T,U)$ defines a parastrophy between $\mathcal{S}_1$ and $\mathcal{S}_2$ then $T^{-1}U\in \mathcal{S}_2$ and $T^{-1}G_1T =U^{-1}G_1U =G_2$. In particular, if $G_1=G_2$ then $U,T \in N_{\GL(d,p)}(G_1)$. \end{lemma} \begin{proof} As $\langle \mathcal{S} \rangle=\langle \mathcal{S}^{-1} \rangle$, it suffices to deal with isotopes. Since $1\in \mathcal{S}_1$, we have $T^{-1}U=T^{-1}1U\in \mathcal{S}_2$. Moreover, $T^{-1}\mathcal{S}_1 T=T^{-1}\mathcal{S}_1 U \cdot U^{-1}T =\mathcal{S}_2\,U^{-1}T \subseteq G_2$, which implies $T^{-1}G_1 T=G_2$. The equation $U^{-1}G_1U=G_2$ follows from $T^{-1}U\in G_2$. \end{proof} We will use the above lemma to compute the classification of quasifields up to parastrophy. \begin{proposition} \label{pr:parast} Let $\mathcal{S}, \mathcal{S}'$ be spread sets of matrices in $\GL(d,p)$. Assume that $1\in \mathcal{S}, \mathcal{S}'$ and $G=\langle \mathcal{S} \rangle = \langle \mathcal{S}' \rangle$. Let $G^*,G^{**}$ be the permutation groups acting on $G$, where the respective actions are the right regular action of $G$ on itself, and the action of $N_{\GL(d,p)}(G)$ on $G$ by conjugation. Let $\iota$ be the inverse map $g\mapsto g^{-1}$ on $G$. Define the permutation group $G^\sharp=\langle G^*, G^{**},\iota \rangle$ acting on $G$. Then, $\mathcal{S}, \mathcal{S}'$ are parastrophic if and only if they lie in the same $G^\sharp$-orbit. \end{proposition} \begin{proof} For any $U\in G$, $T\in N_{\GL(d,p)}(G)$, the sets $T^{-1}\mathcal{S}T$, $\mathcal{S}U$ and $\mathcal{S}^{-1}$ are parastrophes of $\mathcal{S}$. Hence, if $\mathcal{S}, \mathcal{S}'$ are in the same $G^\sharp$-orbit then they are parastrophes. Conversely, assume that the pair $(T,U)$ defines an isotopy between $\mathcal{S}$ and $\mathcal{S}'$. By Lemma \ref{lm:normal}, $\mathcal{S}'=T^{-1}\mathcal{S}T\cdot T^{-1}U$ where $T\in N_{\GL(d,p)}(G)$ and $T^{-1}U\in G$, that is, they are in the same $G^\sharp$-orbit. The proof goes similarly for the case of parastrophy. \end{proof} In general, the explicit computation of the collineation group of a translation plane is very challenging. Another application of Lemma \ref{lm:normal} is the computation of the autotopism group of a quasifield, that is, the computation of the stabilizer of two infinite points in the full collineation group of the corresponding translation plane. \begin{proposition} \label{pr:autotgr} Let $\mathcal{S}$ be a spread set of matrices in $\GL(d,p)$ with $1\in \mathcal{S}$ and $G=\langle \mathcal{S} \rangle$. Define the group \[H=\{(T,U) \in N_{\GL(d,p)}(G)^2 \mid T^{-1}U\in G\}\] and the permutation action $\Phi:G\times H\to G$ of $H$ on $G$ by \[\Phi:(X,(T,U)) \mapsto T^{-1}XU.\] Then, $H$ and $\Phi$ are well defined. The isotopism group of $\mathcal{S}$ is the setwise stabilizer of $\mathcal{S}$ in $H$ with respect to the action $\Phi$. \end{proposition} \begin{proof} In order to see that $H$ is a group, take elements $(T,U), (T_1,U_1)\in H$. On the one hand, $(T^{-1})^{-1}U^{-1}=T(T^{-1}U)^{-1}T^{-1} \in TGT^{-1}=G$, which implies $(T^{-1},U^{-1}) \in H$. On the other hand, $(TT_1,UU_1)\in H$ follows from \[(TT_1)^{-1}UU_1 = T_1^{-1}(T^{-1}U)T_1 \cdot T_1^{-1}U_1 \in T_1^{-1}GT_1\cdot G =G.\] Since $T^{-1}XU=T^{-1}XT\cdot T^{-1}U\in G$ holds for all $X \in G$, $H$ and $\Phi$ are well defined. The claim for the autotopism group of $\mathcal{S}$ follows from Lemma \ref{lm:normal}. \end{proof} As for an exceptional finite transitive linear group $G$, $N_{\GL(d,p)}(G)$ is also exceptional, it is a small subgroup of $\GL(d,p)$ and the autotopism group of $\mathcal{S}$ is computable by GAP4. Using Proposition \ref{pr:collgrps}, we obtain a straightforward method for computing the stabilizer of two infinite points of our translation planes. However, the stabilizer of some infinite points is not an invariant of the translation plane in general. More precisely, let $\Pi$ be a translation plane with infinite line $\ell_\infty$ and $P=\{a_1,\ldots,a_t\} \subseteq \ell_\infty$ be a set of infinite points. Denote by $\mathcal{B}_P$ the pointwise stabilizer of $P$ in $\Aut(\Pi)$. In the rest of this section we show that under some circumstances, the structure of $\mathcal{B}_P$ is an invariant of $\Pi$. We start with a lemma on general permutation groups. \begin{lemma} \label{lm:stab} Let $G$ be a group acting on the finite set $X$ (not necessarily faithfully). For any $Y\subseteq X$, we denoty by $F(Y)$ the pointwise stabilizer of $Y$. Let $Y_1,Y_2$ be subsets of $X$ such that $|Y_1|=|Y_2|< |X|/2$, and suppose that $F(Y_i)$ acts transitively on $X\setminus Y_i$, $i=1,2$. Then, there is an element $g\in G$ with $Y_2^g=Y_1$. In particular, the subgroups $F(Y_1), F(Y_2)$ are conjugate in $G$. \end{lemma} \begin{proof} Up to the action of $G$ on the orbit $Y_2^G$, we may assume that $|Y_1\cap Y_2|\geq |Y_1\cap Y_2^g|$ for all $g\in G$. Suppose that $Y_1\neq Y_2$ and take elements $x\in X\setminus(Y_1\cup Y_2)$, $y_1\in Y_1\setminus Y_2$, $y_2\in Y_2\setminus Y_1$. As $F(Y_i)$ acts transitively on $X\setminus Y_i$, there are elements $g_1\in F(Y_1)$, $g_2 \in F(Y_2)$ such that $x^{g_1}=y_2$ and $x^{g_2}=y_1$. Put $h=g_1^{-1}g_2$. Then $h\in F(Y_1\cap Y_2)$ and $y_2^h=y_1\in Y_1$. This implies $|Y_1\cap Y_2^h|>|Y_1\cap Y_2|$, a contradiction. \end{proof} We can apply the lemma for the stabilizer of infinite points of a translation plane. \begin{proposition} \label{pr:infstab} Let $\Pi$ be a translation plane of order $q$, with infinite line $\ell_\infty$. For a subset $P\subseteq \ell_\infty$, let $\mathcal{B}_P$ denote the pointwise stabilizer subgroup in $\Aut(\Pi)$. Fix the integer $t\leq (q+1)/2$ and define the set \[ \mathcal{D}_t=\{P\subseteq \ell_\infty \mid |P|=t \mbox{ and $\mathcal{B}_P$ act transitively on $\ell_\infty \setminus P$} \}. \] Then, $\Aut(\Pi)$ acts transitively on $\mathcal{D}_t$. In particular, the structure of $\mathcal{B}_P$ (as a permutation group on $\Pi\cup \ell_\infty$) does not depend on the particular choice $P\in \mathcal{D}_t$. \qed \end{proposition} \section{Sharply transitive sets and permutation graphs} Let $G$ be a permutation group acting on the finite set $\Omega$, $n=|\Omega|$ is the degree of $G$. The subset $S\subseteq G$ is a \textit{sharply transitive set of permutations} if for any $x,y\in \Omega$ there is a unique element $\sigma\in S$ such that $x\sigma=y$. Sharply transitive sets can be characterized by the property $|S|=|\Omega|$ and, for all $\sigma,\tau \in S$, $\sigma \tau^{-1}$ is fixed point free. In particular, if $1\in S$ then all $\sigma \in S\setminus \{1\}$ are fixed point free elements of $G$. \begin{definition} Let $G$ be a finite permutation group. The pair $\mathcal{G}=(V,\mathcal{E})$ is the \emph{permutation graph} of $G$, where $V$ is the set of fixed point free elements of $G$ and the edge set $\mathcal{E}$ consists of the pairs $(x,y)$ where $xy^{-1}\in V$. \end{definition} The set $K$ of vertices is a \textit{$k$-clique} if $|K|=k$ and all elements of $K$ are connected. Sharply transitive subsets of $G$ containing $1$ correspond precisely the $(n-1)$-cliques of the permutation graph of $G$. Assume that the action of $G$ on $\Omega$ is imprimitive and let $\Omega'$ be a nontrivial block of imprimitivity. Let $S$ be a sharply transitive subset of $G$ and define the subset $S'=\{\sigma \in S \mid \Omega' \sigma = \Omega'\}$ of $S$. Then, the restriction of $S'$ to $\Omega'$ is a sharply transitive set on $\Omega'$. We will call $\Sigma'$ the \textit{subclique corresponding to the block $\Omega'$}. For the connection between quasifields and sharply transitive sets of matrices see \cite[Chapter 8]{Handbook}. In our computations, we represent quasifields by the corresponding sharply transitive set of matrices, or, more precisely by the corresponding (maximal) clique of the permutation graph. \section{Finite transitive linear groups} \label{sec:trlingr} In this section, we give an overview on finite transitive linear groups. For more details and references see \cite[XII.7.5]{HuppertBlackburn} or \cite[Theorem 69.7]{Handbook}. Let $p$ be a prime, $V=\mathbb F_p^d$, and $\Gamma=\GL(d,p)$. Let $G\leq \Gamma$ be a subgroup acting transitively on $V^*=V\setminus \{0\}$. Then $G_0\trianglelefteq G \leq N_\Gamma(G_0)$, where we have one of the following possibilities for $G$ and $G_0$: \begin{enumerate} \item[1.a)] $G\leq \mathrm{\Gamma L}(1,p^d)$. In particular, $G$ is solvable. \item[1.b)] $G_0\cong \mathrm{SL}(d/e,p^e)$ with $2\leq e\mid d$. \item[2)] $G_0\cong \mathrm{Sp}(d/e,p^e)$ with $e\mid d$, $d/e$ even. \item[3)] $p=2$, $d=6e>6$, and $G_0$ is isomorphic to the Chevalley group $G_2(2^e)$. If $d=6$ then $G_0$ is isomorphic to $G_2(2)'$. Notice that the Chevalley group $G_2(2)$ is not simple, its commutator subgroup has index two and the isomorphism $G_2(2)'\cong \mathrm{PSU}(3,3)$ holds. \item[4)] There are $27$ exceptional finite transitive linear groups, their structure is listed in Table \ref{tab:ecases}. The information given in the last column can be used to generate the sporadic examples in the computer algebra systems GAP4 \cite{GAP4} in the following way. The split extension of $G$ by the vector group $\mathbb F_p^d$ is $2$-transitive, hence primitive, and can be loaded from the library of primitive groups using the command \texttt{PrimitiveGroup($p^d$,$k$)}. As $59^2>2500$, case (4.i) is not included in this library, but since this group is regular on $V^*$, it will not be interesting from out point of view. We have seven sporadic transitive linear groups which are regular, these are denoted by an asterix. These groups have been found by Dickson and Zassenhaus, and they are also known as the right multiplication groups of the \textit{Zassenhaus nearfields.} \end{enumerate} \begin{table} \begin{tabular}{|l|l|l|l|l|} \hline Case & Cond. on $p$ & Cond. on $d$ & $G_0$ & Primit. id. of $p^d:G$\\ \hline (4.a) & $p=5$ & $d=2$ & $\mathrm{SL}(2,3)$ & $15^*,18,19$\\ (4.b) & $p=7$ & $d=2$ & $\mathrm{SL}(2,3)$ & $25^*,29$\\ (4.c) & $p=11$ & $d=2$ & $\mathrm{SL}(2,3)$ & $39^*,42$\\ (4.d) & $p=23$ & $d=2$ & $\mathrm{SL}(2,3)$ & $59^*$\\ (4.e) & $p=3$ & $d=4$ & $\mathrm{SL}(2,5)$ & $124,126,127,128$\\ (4.f) & $p=11$ & $d=2$ & $\mathrm{SL}(2,5)$ & $56^*,57$\\ (4.g) & $p=19$ & $d=2$ & $\mathrm{SL}(2,5)$ & $86$\\ (4.h) & $p=29$ & $d=2$ & $\mathrm{SL}(2,5)$ & $106^*,110$\\ (4.i) & $p=59$ & $d=2$ & $\mathrm{SL}(2,5)$ & no id; regular\\ (4.j) & $p=3$ & $d=4$ & $2^{1+4}$ & $71,90,99,129,130$\\ (4.k) & $p=2$ & $d=4$ & $A_6$ & $16,17$\\ (4.l) & $p=2$ & $d=4$ & $A_7$ & $20$\\ (4.m) & $p=3$ & $d=6$ & $\mathrm{SL}(2,13)$ & $396$\\ \hline \end{tabular} \caption{Exceptional finite transitive linear groups} \label{tab:ecases} \end{table} In this paper, without mentioning explicitly, we consider all finite linear groups as a permutation group acting on the nonzero vectors of the corresponding linear space. \section{Non-existence results for finite right quasifields} \label{sec:nonexistence} In this section, let $(Q,+,\cdot)$ be a finite right quasifield of order $p^d$ with prime $p$. Write $G=\RMlt(Q^*)$ for the right multiplication group of $Q$. As in Section \ref{sec:trlingr}, we denote by $G_0$ a characteristic subgroup of $G$. Moreover, we denote by $S$ the set of nontrivial right multiplication maps of $Q$. Then $1\in S$ and $S$ is a sharply transitive set of permutations in $G$. We write $\mathcal{G}=(V,\mathcal{E})$ for the permutation graph of $G$. Remember that $|S|=p^d-1$ and $S\setminus\{1\}$ is a clique of size $p^d-2$ in $\mathcal{G}$. \begin{proposition} \label{prop:infcl} Assume that $G$ is a transitive linear group belonging to the infinite classes (1)-(3). Then one of the following holds: \begin{enumerate} \item $G\leq \mathrm{\Gamma{}L}(1,p^d)$. \item $G \triangleright \mathrm{SL}(d/e,p^e)$ for some divisor $e<d$ of $d$. \item $p$ is odd and $G \triangleright \mathrm{Sp}(d/e,p^e)$ for some divisor $e$ of $d$. \end{enumerate} \end{proposition} \begin{proof} We have to show that if $p=2$ then the cases $G_0\cong \mathrm{Sp}(d/e,p^e)$ and $G_0\cong G_2(2^e)'$ are not possible. Recall that the transitive linear group $G_2(2^e)$ is a subgroup of $\mathrm{Sp}(6,2^e)$. The impossibility of both cases follow from \cite[Theorem 1]{MuellerNagy}. \end{proof} \begin{proposition} \label{prop:except} If $G$ is an exceptional finite transitive linear group, then it is either a regular linear group or one of those in Table \ref{tab:easrmlt}. \end{proposition} \begin{table} \begin{tabular}{|l|l|l|l|l|} \hline Case & $p^d$ & $G_0$ & Orders (primitive id.)\\ \hline (4.a) & $5^2$ & $\mathrm{SL}(2,3)$ & 48 (18), \textit{96 (19)}\\ (4.b) & $7^2$ & $\mathrm{SL}(2,3)$ & 144 (29)\\ (4.c) & $11^2$ & $\mathrm{SL}(2,3)$ & 240 (42)\\ (4.e) & $3^4$ & $\mathrm{SL}(2,5)$ & 960 (128)\\ (4.f) & $11^2$ & $\mathrm{SL}(2,5)$ & \textit{600 (57)}\\ (4.g) & $19^2$ & $\mathrm{SL}(2,5)$ & 1080 (86)\\ (4.h) & $29^2$ & $\mathrm{SL}(2,5)$ & 1680 (110)\\ (4.l) & $2^4$ & $A_7$ & 2520 (20)\\ \hline \end{tabular} \caption{Exceptional transitive linear groups as right multiplication groups} \label{tab:easrmlt} \end{table} This proposition is proved in several steps. \begin{claim} \label{claim:no_4k4m} $G$ cannot be an exceptional transitive linear group of type (4.k) and (4.m). \end{claim} \begin{proof}[Computer proof] Assume $G$ of type (4.k) and $G_0\cong A_6$. Let $S$ be a Sylow $5$-group of $G$. Let $A,B,C$ the orbits of $S$ such that $B\cup C$ is an orbit of $N_G(S)$. Then, the sets $A,B$ of size $5$ satisfy $|A\cap B^g|\in\{0,2\}$ for all $g\in G$. By \cite[Lemma 2]{MuellerNagy}, $G$ does not contain a sharply transitive set of permutations. Assume $G$ of type (4.m). Let $S$ be a Sylow $7$-group of $G$ and $H=N_G(S)$. Then $|H|=28$ and $H$ has orbits of length $28$. One can find $H$-orbits $A,B$ of size $28$ such that $|A\cap B^g|\in\{0,6\}$ for all $g\in G$. Again by \cite[Lemma 2]{MuellerNagy}, $G$ does not contain a sharply transitive set of permutations. \end{proof} \begin{claim} \label{claim:no_4e} $G$ cannot be an exceptional transitive linear group of type (4.e) and order $240$ or $480$. \end{claim} \begin{proof}[Computer proof] We proceed similarly to the cases in \ref{claim:no_4k4m}. Assume $G$ of type (4.e) and $|G| \in \{240,480\}$. Let $H$ be the normalizer of a Sylow $5$-group in $G$. Then $|H|=40$ and $H$ has orbits $A,B$ of length $40$ such that $|A\cap B^g|\in\{0,24\}$ for all $g\in G$. This proves the claim by \cite[Lemma 2]{MuellerNagy} with $p=3$. \end{proof} Beside finite regular linear groups, the remaining exceptional transitive linear groups are those of type (4.j), that is, when $G_0$ is the extraspecial group $E_{32}^+$ of order $2^5$. In this class, there are five groups, three of them are solvable. The cases of the three solvable group of type (4.j) were left unsolved by M. Kallaher \cite{Kallaher}, as well. The computation is indeed very tedious even with today's hardware and software. The construction of the groups of type (4.j) is as follows. $\GL(4,3)$ has a unique subgroup $L_0$ of order $2^5$ which is isomorphic to the extraspecial $2$-group $E_{32}^+$. The normalizer $L=N_{\GL(4,3)}(L_0)$ has order $3840$. $L$ has five transitive linear subgroups containing $L_0$; the orders are $160, 320, 640, 1920, 3840$. Clearly, it suffices to prove the nonexistence of $79$-cliques for the permutation graph of $L$ only. The action of $L$ is not primitive; it has blocks of imprimitivity of size $2,8$ and $16$. These blocks are unique up to the action of $L$. \begin{claim} \label{claim:non2gr} Let $A$ be a block of $L$ of size $16$. Let $K$ be a $15$-clique corresponding to $A$ and assume that the group $\langle K \rangle$ generated by $K$ is not a $2$-group. Then, $K$ cannot be extended to a $79$-clique of $L$. \end{claim} \begin{proof}[Computer proof] Let $A$ be a block of size $8$. Denote by $H$ the setwise stabilizer of $A$ in $L$; $H$ stabilizes another block $B$ of size $8$ and $A\cup B$ is a block of size $16$. Let $\mathcal{A}$ be the set of all $7$-cliques corresponding to the block $A$. As $N_L(H)$ operates on the permutation (sub)graph of $H$, it also acts on $\mathcal{A}$; let $\mathcal{A}_0$ be a set of orbit representatives. We use \cite{Grape} to compute $\mathcal{A}_0$; $|\mathcal{A}_0|=98$. Then in all possible ways, we extend the elements of $\mathcal{A}_0$ to $15$-cliques corresponding to the block $A\cup B$, let $\mathcal{B}$ denote the set of extended cliques. Finally, we filter out the $15$-cliques generating a non-$2$-group and show that none of them can be extended to a $79$-clique. \end{proof} We have to deal with subcliques generating a $2$-group. An important special case is the following. \begin{claim} \label{claim:15clique} Up to conjugacy in $L$, there is a unique $15$-clique $K^*$ corresponding to an imprimitivity block of size $16$ such that the setwise stabilizer in $L$ has order $192$. $K^*$ has the further properties: \begin{enumerate} \item The subgroup $\langle K^* \rangle$ has order $32$. \item $\langle K^* \rangle$ interchanges two blocks of size $16$ and fixes the other three. \end{enumerate} \end{claim} \begin{proof}[Computer proof] Let $S$ be a Sylow $2$-subgroup of $L$. Using \cite{Grape}, one can compute all $15$-cliques of the permutation graph of $S$. Up to conjugacy in $S$, there are $17923$ such cliques, only one of them has stabilizer of size $192$. The properties of $K^*$ are obtained by computer calculations. \end{proof} \begin{claim} \label{claim:2gr} Let $A$ be a block of $L$ of size $16$. Let $K$ be a $15$-clique corresponding to $A$ and assume that the group $\langle K \rangle$ generated by $K$ is a $2$-group. Then, $K$ cannot be extended to a $79$-clique of $L$. \end{claim} \begin{proof}[Computer proof] Let $S$ be a Sylow $2$-subgroup of $L$. $S$ leaves a block of size $16$, say $A$, invariant. We compute the set $\mathcal A$ of $S$-orbit representatives of the $15$-cliques of the permutation graph of $S$. $\mathcal{A}$ contains precisely one element conjugate to $K^*$, in fact, we assume that $K^*\in\mathcal{A}$. For all elements $K \in \mathcal{A}\setminus \{K^*\}$, the computer shows within a few seconds that $K$ cannot be extended to a $79$-clique. For $K^*$, the direct computation takes too long, we therefore give a theoretical proof. Let us assume that $D$ is a clique of size $79$ in the permutation graph of $L$. We may assume that all subcliques of $D$ corresponding to $16$-blocks are conjugate of $K^*$. Denote by $D_A$ the subclique of $D$ of size $15$, corresponding to $A$. By \ref{claim:15clique}, $\langle D_A\rangle$ interchanges two $16$-blocks, say $B,B'$, and leaves the others invariant. Denote by $D_B$ the subclique of $D$, corresponding to $B$. Clearly, $D_A\neq D_B$. As $\langle D_B \rangle$ leaves three blocks invariant, there is a $16$-block $C$ which is invariant under all elements of $D_A \cup D_B$. However, as $|C|=16$, $D$ cannot have more than $15$ elements mapping $C$ to $C$, a contradiction to $|D_A\cup D_B|>15$. \end{proof} The claims \ref{claim:non2gr}, \ref{claim:15clique} and \ref{claim:2gr} imply the following result. \begin{claim} \label{claim:noextraspec} $G$ cannot be an exceptional transitive linear group of type (4.j). \end{claim} The combination of the claims \ref{claim:no_4k4m}, \ref{claim:no_4e} and \ref{claim:noextraspec} yields the proof of the Proposition \ref{prop:except}. \section{Exhaustive search for cliques and their invariants} \label{sec:exhaustive} Let $G\leq \GL(d,p)$ be a transitive linear group. We use the program CLIQUER \cite{Cliquer} to compute all cliques of size $p^d-1$ in the permutation graph $\mathcal{G}$ of $G$. For the exceptional transitive linear groups of Table \ref{tab:ecases}, the result is presented in the second column of Table \ref{tab:maxcls}. Proposition \ref{pr:parast} allows us to reduce our results on cliques in exceptional transitive linear groups modulo parastrophy, as shown in the third column of Table \ref{tab:maxcls}. In column 4, we filtered out those cliques which do not generate the whole group $G$. \begin{table} \begin{tabular}{|l||c|c|c|c|} \hline type & \# of cliques & up to parastrophy & proper $G$ & \# CCFPs'\\ \hline\hline (4.a) & $4;8$ & $2;3$ & $1;0$ & $1;0$ \\ \hline (4.b) & 12 & 4 & 2 & 2 \\ \hline (4.c) & 16 & 4 & 3 & 3 \\ \hline (4.e) & 27648 & 32 & 21 & 20 \\ \hline (4.f) & 6 & 2 & 0 & 0 \\ \hline (4.g) & 9 & 3 & 3 & 3 \\ \hline (4.h) & 64 & 9 & 8 & 8 \\ \hline (4.l) & 450 & 2 & 2 & 1 \\ \hline \end{tabular} \caption{Maximal cliques in exceptional transitive linear groups} \label{tab:maxcls} \end{table} In the final step, we compute the \textit{Conway-Charnes fingerprint} of all spread sets of matrices, cf. \cite{CharnesDempwolff,MathonRoyle}. The Conway-Charnes fingerprint is an invariant of the translation plane which can be easily computed from any spread set of matrices. \begin{definition} Given a spread set $\mathcal{S}=\{U_1,\ldots,U_{q-1}\}$ of nonzero matrices, one forms a $(q-1) \times (q-1)$ matrix whose $(i,j)$ entry is $0$, $+1$ or $-1$ according as $\det(U_i-U_j)$ is zero, square or nonsquare, respectively. Border this matrix with a leading row and column of 1s (except for the first entry on the diagonal which remains zero) to form a symmetric $q\times q$ matrix $A$ with $0$ on the diagonal, and $\pm 1$ in every non-diagonal entry. Finally, form the matrix $F = AA^t$ (with the product being taken in the rational numbers). The \emph{fingerprint} of the spread set is the multiset of the absolute values of the entries of $F$. \end{definition} The last column of Table \ref{tab:maxcls} contains the number of different Conway-Charnes fingerprints of the spread sets of matrices in the corresponding exceptional transitive linear group $G$. One sees that only for two pairs of spread sets do we obtain the same fingerprint. Let us denote by $Q_1,Q_1'$ the quasifields of order $3^4$ and by $Q_2,Q_2'$ those of order $2^4$. We compute the corresponding autotopism groups $\mathcal{A}_1, \mathcal{A}_1'$ and $\mathcal{A}_2,\mathcal{A}_2'$. GAP4 shows that $\mathcal{A}_1$ and $\mathcal{A}_1'$ are nonisomorphic groups of order $640$, acting transitively on the $80$ points of $\ell_\infty \setminus \{(0), (\infty)\}$. By Proposition \ref{pr:infstab}, $Q_1,Q_1'$ determine nonisomorphic translation planes. $\mathcal{A}_2$ and $\mathcal{A}_2'$ are both isomorphic to $\mathrm{PSL}(2,7)$. Both groups fix a third point $a,a'$ of the infinite line and act transitively on the remaining $14$ infinite points. Hence, both groups are the stabilizer of a triple of infinite points and Proposition \ref{pr:infstab} applies. As the orbit lengths of $\mathcal{A}_2$ and $\mathcal{A}_2'$ are different, we can conclude that the two translation planes are nonisomorphic. \section{Right multiplication groups of finite right quasifields} We compile our results in the following theorem. \begin{theorem} Let $(Q,+,\cdot)$ be a finite right quasifield of order $p^d$ with prime $p$. Then, for $G=\RMlt(Q^*)$, one of the following holds: \begin{enumerate} \item $G\leq \mathrm{\Gamma{}L}(1,p^d)$ and the corresponding translation plane is a generalized Andr\'e plane. \item $G \triangleright \mathrm{SL}(d/e,p^e)$ for some divisor $e<d$ of $d$ with $e\neq d$. \item $p$ is odd and $G \triangleright \mathrm{Sp}(d/e,p^e)$ for some divisor $e$ of $d$. \item $p^d\in \{5^2,7^2,11^2,17^2,23^2,29^2,59^2\}$ and $G$ is one of the seven finite sharply transitive linear groups of Zassenhaus \cite{Zassenhaus}. The corresponding translation planes are called Zassenhaus nearfield planes. \item $p^d\in \{5^2,7^2,11^2\}$, and $G$ is a solvable exceptional transitive linear group. These quasifields and the corresponding translation planes have been given by M. J. Kallaher \cite{Kallaher}. \item $p^d=3^4,19^2$ or $29^2$, and the number of translation planes is $21$, $3$ or $8$, respectively. \item $p^d=16$ and $G=A_7$. The corresponding translation planes are the Lorimer-Rahilly and Johnson-Walker planes. \end{enumerate} \end{theorem} \begin{proof} By Proposition \ref{prop:infcl}, (1), (2) or (3) holds if $G$ belongs to one of the infinite classes of finite transitive linear groups. If $G\leq \mathrm{\Gamma{}L}(1,p^d)$ then the corresponding translation plane is a generalized Andr\'e plane by \cite[Theorem 3.1]{Kallaher}. The cases of the Zassenhaus nearfield planes are in (4). The arguments in Sections \ref{sec:nonexistence} and \ref{sec:exhaustive} imply (6) and that there are two quasifields having $A_7$ as right multiplication group. Moreover, the corresponding translation planes are nonisomorphic. \cite[Corollary 4.2.1]{JhaKallaher} implies that the two translation planes are the Lorimer-Rahilly and Johnson-Walker planes. \end{proof} We close this paper with two remarks. \begin{enumerate} \item No quasifield $Q$ is known to the author with $\RMlt(Q^*) \triangleright \mathrm{Sp}(d/e,p^e)$. Even the smallest case of $\mathrm{Sp}(4,3)$ is computationally challenging. \item The Lorimer-Rahilly and Johnson-Walker translation planes are known to be polar to each other, that is, one spread set of matrices is obtained by transposing the matrices in the other spread set. (See \cite[29.4.4]{Handbook}.) \end{enumerate} The GAP programs used in this paper are avaible on the author's web page: \begin{center} \verb+http://www.math.u-szeged.hu/~nagyg/pub/rightmlt.html+ \end{center} \bibliographystyle{plain}
\section{Introduction} In the classical theory of determinantal hypersurfaces, the case of pfaffian cubics has already found many applications. (\cite{Ma-Ti}, \cite{I-R}, \cite{Dr}). This paper presents new invariants for these objects. Let ${\mathbb{P}_5}$ be a five dimensional projective space over the complex numbers, and denote by $V_6$ the vector space $H^0 ({\cal O}_{{\mathbb{P}_5}}(1))$. For $n\geq 2$, let $\pi_{n-1}$ be a projective space of dimension $n-1$, and $W_n=H^0({\cal O}_{\pi_{n-1}}(1))$. \begin{definition}\label{defpfaf} For $2\leq n$, a general element of $\bigwedge\limits^2 V_6 \otimes W_n$ gives a skew-symmetric map $M$ with linear coefficients over $\pi_{n-1}$. We have the exact sequence: $$\begin{CD} 0@>>> V_6^\vee \otimes {\cal O}_{\pi_{n-1}}(-1) @>M>> V_6 \otimes {\cal O}_{\pi_{n-1}} @>>> F@>>> 0 \end{CD}$$ where $F$ is a rank $2$ sheaf over the pfaffian divisor. For $n\leq 6$, the sheaf $F$ is a vector bundle over a smooth cubic. Let's call it the pfaffian bundle defined by $M$. \end{definition} It is known from classical works on representations of a cubic form by pfaffians (Cf \cite{Be}, \cite{Do}), that for $n\geq 6$ a general cubic divisor is not a pfaffian, and for $3\leq n \leq 5$ the pfaffian bundles have moduli spaces of strictly positive dimension. \bigskip The main result of this article concerns the $n=4$ case. In this situation we have the following results from \cite{Be}: \begin{itemize} \item[-] Every smooth cubic surface of $\mathop{\pi_3}$ can be defined by a linear pfaffian. \item[-] Let $(W_4\otimes \bigwedge\limits^2 V_6)^{sm}$ be the openset of $W_4\otimes \bigwedge\limits^2 V_6$ corresponding to smooth pfaffian surfaces. For any element $M$ of $(W_4\otimes \bigwedge\limits^2 V_6)^{sm}$, the pfaffian sheaf is a stable rank $2$ vector bundle over the pfaffian surface $Pf(M)$. Moreover, it is an arithmetically Cohen-Macaulay sheaf, and every arithmetically Cohen-Macaulay rank $2$ vector bundle over a smooth cubic surface $S$ with determinant ${\cal O}_S(2)$ is a pfaffian bundle. \item[-] The quotient of $(W_4\otimes \bigwedge\limits^2 V_6)^{sm}$ by $GL(V_6)$ with the following action: $$\begin{array}[c]{ccc} GL(V_6)\times (W_4\otimes \bigwedge\limits^2 V_6)^{sm} & \to & (W_4\otimes \bigwedge\limits^2 V_6)^{sm} \\ (P,M) & \mapsto & ^t P.M.P \end{array}.$$ is isomorphic to the space of pairs $(S,F)$ where $S$ is a smooth cubic surface of $\mathop{\pi_3}$ and $F$ an isomorphism class of a pfaffian bundle on $S$. It is a geometric quotient. \end{itemize} In this article we obtain a geometric interpretation of these orbits. \begin{definition}\label{inscribed} A complete pentahedron inscribed in a cubic surface $S$ of the projective space $\mathop{\pi_3} $ is a set $\{H_0, \dots, H_4\}$ of $5$ planes of $\mathop{\pi_3} $ such that: \begin{itemize} \item[i)] $(H_0, \dots, H_4)$ is a projective basis of $\mathop{\pi_3 ^\vee}$. \item[ii)] The $10$ points $(H_i\cap H_j \cap H_k)_{0\leq i<j<k\leq 4}$ are on $S$. \end{itemize} We define the subset $\mathcal{H}$ of $|O_{\mathop{\pi_3}}(3)|\times |O_{\mathop{\pi_3}}(5)|$ (resp. the subset $\mathcal{H}_{ord}$ of $|O_{\mathop{\pi_3}}(3)|\times (\mathop{\pi_3})^5$) to be the set of elements $(S,\Pi)$ such that $S$ is a smooth cubic surface of $\mathop{\pi_3}$ and $\Pi$ is a complete pentahedron inscribed in $S$ (resp. complete pentahedron inscribed in $S$ with an order on the five planes). For a cubic surface $S$ of $\mathop{\pi_3}$, denote by $\mathcal{H}_S$ the pullback in $\mathcal{H}$ of $S$ by the first projection. \end{definition} The first four sections give two natural methods to construct $5$ hyperplane sections of a cubic surface from a pfaffian vector bundle. Eventually they are generically identical, and we obtain: \newpage \begin{theorem}\label{mainbir} There is a natural birational map from $(W_4\otimes \bigwedge\limits^2 V_6)^{sm}/GL(V_6)$ to $\mathcal{H}$ such that the composition with the projection to $|{\cal O}_{\mathop{\pi_3}}(3)|$ is the pfaffian map. In particular: \begin{itemize} \item[.] $(W_4\otimes \bigwedge\limits^2 V_6)^{sm}/GL(V_6)$ is a rational variety of dimension $24$. \item[.] Let $S$ be a general cubic surface. The moduli space of pfaffian bundles on $S$ is birational to $\mathcal{H}_S$. \end{itemize} \end{theorem} Both constructions enlight this theorem differently. The first one: $\Phi_1$ (cf definition \ref{defPHI1}) is a classical problem of hyperplane restriction of the pfaffian bundles. So section \ref{secn3} starts with the easy case $n=3$ to introduce some invariants of these bundles. The universal situation is then described because many geometric interpretations of the later sections are specializations of this construction. \bigskip Section \ref{secn4} details the $n=4$ case to settle the construction of $\Phi_1$. The projectivisation of a pfaffian bundle on a cubic surface is called a Palatini threefold. Such varieties are well-known to be the only known examples of smooth $3$-dimensional varieties $X$ in ${\mathbb{P}_5}$ such that $h^0({\cal O}_X(2))>h^0({\cal O}_{{\mathbb{P}_5}}(2))$. One can find references for their Hilbert scheme (\cite{Fa-Fa}, \cite{Fa-Me}), and also references where they are in a list of exceptions to some geometrical property (cf \cite{Me-Po}, \cite{Ot}). Some of their classical properties are also described in \cite{Do} and \cite{Ok}. But here some new results are required. First we give an interpretation of their anticanonical linear system to prove that it is $\mathop{\pi_3 ^\vee}$. Then we describe this linear system in proposition \ref{propantican}. It turns out that its exceptional locus is $5$ points of $\mathop{\pi_3 ^\vee}$. This achieves the construction of $\Phi_1$. \bigskip But the geometric configuration of these planes is only explained in section \ref{W2V6} by the second construction: $\Phi_2$ (cf corollary \ref{defPHI2}). This time, it is a problem of sum of matrices of rank $2$. The key step to construct $\Phi_2$ is the surprising proposition \ref{P4-5sec} with following summary: \begin{proposition-non} The projection of the Grassmannian of lines $G(2,V_6)$ from a general $3$-dimensional projective space has a single point of order $5$. \end{proposition-non} The claim that $\Phi_1$ and $\Phi_2$ are generically the same, and also their birationality, are proved in section \ref{sectionexplicit} from the explicit formula of theorem\ref{explicit}. This ends the proof of theorem \ref{mainbir}. As a by-product we obtain in corollary \ref{unirationalisation} an explicit generically finite unirationalization of the quotient of the product of five copy of $G(2,V_6)$ by the diagonal action of $PGL(V_6)$. \smallskip Recently, F. Tanturri (cf \cite{T}) found an algorithm to obtain a pfaffian representation from the equation of a cubic surface. Although some representations are similar, the main difference is that any pfaffian bundle on the surface would solve his problem, while in our situation we have additional requirements such that only one bundle is solution. \bigskip In the last section we investigate those properties over a base. We explain how the Debarre-Voisin's symplectic manifold can be considered as a parameter space for Palatini threefolds in a six dimensional variety of $\mathbb{P}_9$. Those varieties of dimension six were discovered by C. Peskine. They are of independent interest because they are smooth and non quadratically normal in $\mathbb{P}_9$ (it's a boundary case in Zak's theory of quadratic normality). However, most of their geometric properties are unknown. In particular, it would be very interesting to understand those varieties from a Palatini threefold in a similar way that a Veronese surface is related to $\mathbb{P}_2\times \mathbb{P}_2$. So we will also explain in this section the consequences on the Peskine's varieties of the work on the Palatini threefolds done in section \ref{secn4}. \begin{description} \item[\bf Aknowledgement:] \ \\ I'd like to thanks I. Dolgachev for encouraging discussions and references. \end{description} \section{Invariants of Pfaffian bundles over plane cubics.}\label{secn3} \subsection{Ruled surfaces in ${\mathbb{P}_5}$, and the $n=3$ case.} In this section, we detail the case $n=3$. The following easy lemma is a basic step that enlights the next sections. \begin{lemma}\label{planecubi} For a general element of $W_3\otimes \bigwedge\limits^2 V_6$, we consider the associated exact sequence: \begin{equation}\label{pfaffP2} \begin{CD} 0@>>> V_6^\vee \otimes {\cal O}_{\pi_{2}}(-1) @>M>> V_6 \otimes {\cal O}_{\pi_{2}} @>>> F@>>> 0 \end{CD} \end{equation} with $M= -\t{M}$. The cokernel $F$ is a rank $2$ vector bundle over the smooth plane cubic $C$ defined by the pfaffian of $M$, and $F$ is isomorphic to one of the following bundles: \begin{itemize} \item[a)] ${\cal L}(1) \oplus {\cal L} ^{\vee}(1)$, where ${\cal L}$ is a line bundle of degree $0$ on $C$ such that $h^0({\cal L}^2)=0$. \item[b)] $F$ is the unique unsplit extension: $$0 \rightarrow \theta(1) \rightarrow F\rightarrow \theta(1)\rightarrow 0 $$ where $\theta^2= {\cal O}_C$ and $\theta\neq {\cal O}_C$. \item[c)] $F=\theta(1) \oplus \theta(1)$ where $\theta^2= {\cal O}_C$ and $\theta\neq {\cal O}_C$. \end{itemize} \end{lemma} {\noindent\it Proof: } To simplify the notations, let $F_0$ denote $F(-1)$. First one can remark that $h^0(F_0)=0$, and that $F_0 \simeq (F_0) ^{\vee}$ because $M$ is skew-symmetric. So we have $\wedge^2(F_0) = {\cal O}_C $. We choose a point $p$ on $C$. We will now prove that there is a point $r$ of $C$ such that $h^0(F_0(p-r))>0$. From Riemann-Roch's theorem the bundle $F_0(p)$ has a pencil of sections. This gives, on $\mathbb{P}_1\times C$, a section of the bundle ${\cal O}_{\mathbb{P}_1}(1) \boxtimes F_0(p)$. But the computation of the second Chern's class of this bundle implies that this section has a non empty vanishing locus, so there is a point $r$ of $C$ such that $h^0(F_0(p-r))>0$. Let's recall that $h^0(F_0)=0$ to obtain that ${\cal O}_C(p-r)$ is not trivial and that $F$ is isomorphic to one of the $3$ above cases. \nolinebreak $\Box$ \begin{remark}\label{3casplans} The ruled surface $\Proj{F}$ has a natural embedding in ${\mathbb{P}_5}$ given by the surjection in the sequence (\ref{pfaffP2}) such that in the cases: \begin{itemize} \item[a)] it contains $2$ plane cubic curves, and the planes spanned by these curves are disjoint in ${\mathbb{P}_5}$. \item[b)] it contains only one plane cubic. \item[c)] it contains infinitely many plane cubics. The planes spanned by these curves are the planes of a Segre: $\mathbb{P}_1\times \mathbb{P}_2 \subset {\mathbb{P}_5}$. \end{itemize} Moreover, the planes in those $3$ cases are the planes of ${\mathbb{P}_5}$ isotropic for all the skew-symmetric forms defined by $M$. \end{remark} {\noindent\it Proof: } In those $3$ cases, the bundle $F$ has an invertible quotient of rank $1$ and degree $3$. We just have to show that those embeddings of $C$ are isotropic for $M$. But it is a corollary of the fact that the resolution of $F$ can have a skew-symmetric form deduced from the isomorphism: $\wedge^2(F(-1)) \simeq {\cal O}_C$. Conversely, any isotropic plane for $M$ gives the existence of $P\in GL(V_6)$ such that: $\t{P}.M.P= \left(\begin{array}{cc} 0 & -\t{A}\\ A & B \end{array}\right) $, where $A,B$ are $3$ by $3$ matrices with linear entries. So the cokernel of $A$ gives the expected invertible quotient of $F$ of degree $3$. \nolinebreak $\Box$ \subsection{Universal settings and the $SL(V_6)$-invariant double cover} \begin{definition} Let $G(3,V_6 ^{\vee})$ and $G(3,\bigwedge\limits^2 V_6)$ be the Grassmannians of $3$-dimensional vector subspaces of $V_6 ^{\vee}$ and $\bigwedge\limits^2 V_6$. Denote by $K_3$ and $R_3$ their tautological subbundles. We define the isotropic incidence: $$ \begin{CD} Z\subset G(3,V_6 ^{\vee}) \times G(3,\bigwedge\limits^2 V_6) @>p_2>> G(3,\bigwedge\limits^2 V_6)\\ @Vp_1VV \\ G(3,V_6 ^{\vee}) \end{CD} $$ to be the vanishing locus of the unique $SL(V_6)$-invariant section of $\bigwedge\limits^2 K_3 ^{\vee} \boxtimes R_3 ^{\vee}$. Denote by $\mathcal{U}$ the open subset of $G(3, \bigwedge\limits^2 V_6)$ made of subspaces such that the intersection of their projectivisation with the pfaffian hypersurface of $\mathbb{P}(\bigwedge\limits^2 V_6)$ is a smooth cubic curve. The restriction of $Z$ to $G(3,V_6 ^{\vee})\times \mathcal{U}$ will be noted: $Z_{\mathcal{U}}$. Let $E_{12}$ be the rank $12$ bundle defined by the exact sequence: \begin{equation} \label{eqE12} \begin{CD} 0@>>> E_{12} @>>> \bigwedge\limits^2 V_6 \otimes {\cal O}_{G(3,V_6 ^{\vee})} @>>> \bigwedge\limits^2 K_3 ^{\vee} @>>> 0 \end{CD} \end{equation} \end{definition} I'd like to thanks A. Kuznetsov for the following description of $Z$ from the relative Grassmannian. \begin{proposition}\label{propisoE12} The isotropic incidence $Z$ is isomorphic to the relative Grassmannian $G(3,E_{12})$ of linear subspaces of the bundle $E_{12}$. The projection $Z_{\mathcal{U}} \to \mathcal{U}\subset G(3,\bigwedge\limits^2 V_6)$ is generically finite of degree $2$. The fibers of this morphism over an element of type a,b,c in Lemma \ref{planecubi} is respectively in $G(3,V_6 ^{\vee})$: $2$ points, $1$ point, and a rational cubic curve. \end{proposition} {\noindent\it Proof: } Let $(\mu,\nu)$ be an element of $G(3,V_6 ^{\vee})\times G(3,\bigwedge\limits^2 V_6)$. The fiber of a vector bundle at $\mu$ (resp. $\nu$) will be noted by the name of the bundle with the index $\mu$ (resp. $\nu$). The vector space $K_{3,\mu}$ is isotropic for all the skew-symmetric forms defined by the elements of $R_{3,\nu}$ if and only if $(\mu,\nu)\in Z$, but also if and only if the composition: $$ \begin{CD} R_{3,\nu} @>>> \bigwedge\limits^2 V_6 @>>> \bigwedge\limits^2 K_{3,\mu} ^{\vee} \end{CD} $$ is the zero map. So $(\mu,\nu)\in Z \iff R_{3,\nu} \subset E_{12,\mu}$ and we have the equality $Z=G(3,E_{12})$. The end of the assertion follows immediatly from Lemma \ref{planecubi} and Remark \ref{3casplans}. \nolinebreak $\Box$ \begin{corollary} The locus $\mathcal{U}_c$ in $\mathcal{U}\subset G(3,\bigwedge\limits^2 V_6)$ of planes of type c) has codimension $3$. Consider the following relation on $\mathcal{U}_c$: $ p \mathcal{R} p'$ if and only if $p_1(p_2^{-1}(p))=p_1(p_2^{-1}(p'))$. For any element $p$ of $ \mathcal{U}_c$, there is a six dimensional subspace $L_p$ of $\bigwedge\limits^2 V_6$ such that the equivalence class of $p$ for $\mathcal{R}$ is an open set of $G(3,L_p)$. \end{corollary} {\noindent\it Proof: } From the proposition \ref{propisoE12}, for any $p$ in $\mathcal{U}_c$, $p_1(p_2^{-1}(p))$ is a smooth rational cubic curve $C_p$ in $G(3,V_6 ^{\vee})$. So the restriction of $E_{12}$ to $C_p$ is $6{\cal O}_{\mathbb{P}_1}\oplus 6{\cal O}_{\mathbb{P}_1}(-1)$, and this bundle has a natural trivial subbundle of rank $6$. Let $L_p$ be the six dimensional vector subspace of $\bigwedge\limits^2 V_6$ obtained from the image of this subbundle by the injection of the sequence (\ref{eqE12}). Proposition \ref{propisoE12} describes $p_1^{-1}(C_p)$ as the relative Grassmannian $G(3,E_{12|C_p})$. Let $F$ be a subvector bundle of rank $3$ of $E_{12|C_p}=L_p\otimes{\cal O}_{\mathbb{P}_1}\oplus 6{\cal O}_{\mathbb{P}_1}(-1)$. The case c) appears when the line bundle $\wedge^3 F ^{\vee}$ contracts the curve $C_p$. But $\wedge^3 F ^{\vee}$ is not ample if and only if $F$ is a trivial subbundle of $L_p\otimes{\cal O}_{\mathbb{P}_1}$. So $p_1^{-1}(C_p)\cap p_2^{-1}(\mathcal{U}_c)$ is $(\mathcal{U}\cap G(3,L_p))\times C_p$, and the equivalence classe of $p$ for $\mathcal{R}$ is $\mathcal{U}\cap G(3,L_p)$. So the dimension of $\mathcal{U}_c$ is the sum of the dimension of $G(3,6)$ with the dimension of the family of rational cubic curves in $G(3,V_6 ^{\vee})$ . In conclusion $\mathcal{U}_c$ has dimension $33$ and codimension $3$ in $G(3,\bigwedge\limits^2 V_6)$. \nolinebreak $\Box$ \section{Palatini threefolds}\label{secn4} In this section we study the case $n=4$. \subsection{D\'efinition and classical properties} \begin{definition}\label{palatini} A smooth $3$ dimensional sub-variety $X$ of ${\mathbb{P}_5}$ is called a Palatini threefold\footnote{or a Palatini scroll} if there exists an element of $\alpha \in \bigwedge\limits^2 V_6 \otimes W_4$ such that $X=\Proj{F}$ where $F$ is the pfaffian vector bundle defined from $\alpha$ in the Definition \ref{defpfaf} with $n=4$. It is also classically called (cf \cite{Do} p589) the singular variety of the linear system $|W_4 ^{\vee}|$ of linear line complex in $|V_6 ^{\vee}|$. \end{definition} \begin{notation}\label{notapala} In this section, denote by $X$ a Palatini threefold in ${\mathbb{P}_5}$, by $h$ the class of an hyperplane of ${\mathbb{P}_5}$, by $S$ the pfaffian cubic surface in $\mathop{\pi_3}$ and by $s$ the pullback on $X$ of the class of an hyperplane of $\mathop{\pi_3}$. The cotangent bundle of ${\mathbb{P}_5}$ will be noted $\Omega^1_{{\mathbb{P}_5}}$. \end{notation} So we can immediately obtain the well known resolution of its ideal: \begin{remark}\label{resolpala} The ideal $I_X$ of a Palatini threefold $X$ in ${\mathbb{P}_5}$ has the following resolution: \begin{equation} \label{resopalaP5} \begin{CD} 0@>>> W_4 ^{\vee} \otimes {\cal O}_{{\mathbb{P}_5}} @>\alpha>> \Omega^1_{{\mathbb{P}_5}}(2h) @>>> I_X(4h) @>>> 0 \end{CD} \end{equation} \end{remark} and the famous equality: $$h^0{\cal O}_X(2h)=h^0{\cal O}_{{\mathbb{P}_5}}(2h)+1.$$ \bigskip To explain the natural embedding of $X$ in the point/plane incidence of ${\mathbb{P}_5}$, F. Zak introduced the following vector bundle: \begin{definition} The canonical extension on ${\mathbb{P}_5}$ displayed in the second column of the following diagram of exact sequences $$ \begin{CD} @. @. 0 @. @.\\ @. @. @VVV @. \\ 0 @>>> W_4 ^{\vee} \otimes {\cal O}_{{\mathbb{P}_5}}(-h) @>>> \Omega^1_{{\mathbb{P}_5}}(h) @>>> I_X(3h) @>>>0\\ @. @V{\sim}VV @VVV \\ 0@>>> W_4 ^{\vee} \otimes {\cal O}_{{\mathbb{P}_5}}(-h) @>\alpha>> V_6\otimes {\cal O}_{{\mathbb{P}_5}} \\ @. @. @VVV @. \\ @. @. {\cal O}_{{\mathbb{P}_5}}(h) @. @.\\ @. @. @VVV @. \\ @. @. 0 @. @.\\ \end{CD} $$ induces on a Palatini threefold $X$ the following extension with middle term a rank $3$ vector bundle $E_X$. $$ \begin{CD} 0 @>>> N_X ^{\vee}(3h) @>>> E_X @>>> {\cal O}_X(h) @>>> 0. \end{CD} $$ Moreover, the restriction to $X$ of the second line of the previous diagram gives the exact sequence: \begin{equation} \label{resEX} \begin{CD} 0@>>> {\cal O}_X(-h-s)@>>> W_4 ^{\vee} \otimes {\cal O}_{X}(-h) @>\alpha>> V_6\otimes {\cal O}_{X} @>>> E_X @>>> 0 \end{CD} \end{equation} and the determinant of $E_X$ is ${\cal O}_X(3h-s)$. \end{definition} \bigskip From the inclusion $W_4 ^{\vee} \subset \bigwedge\limits^2 V_6$ and the identification $W_4=\wedge^3W_4 ^{\vee}$, we can consider $\mathop{\pi_3 ^\vee}$ as a subvariety of $G(3,\bigwedge\limits^2 V_6)$. \begin{proposition}\label{Z4isoX} Let $Z_4$ be the restriction of the isotropic incidence $Z\subset G(3,V_6 ^{\vee})\times G(3,\bigwedge\limits^2 V_6)$ to $G(3,V_6 ^{\vee})\times \mathop{\pi_3 ^\vee}$. Then $Z_4$ is isomorphic to $X$ and the projection from $Z_4$ to $G(3,V_6 ^{\vee})$ is the natural embedding of $X$ given by $\bigwedge\limits^3 E_X$. \end{proposition} {\noindent\it Proof: } Let's first recall the classical description of quadrisecant lines to $X$. Let $A ^{\vee}$ and $B$ be the $3$ dimensional vector subspaces of $V_6 ^{\vee}$ and $W_4 ^{\vee}$ corresponding to a point of $Z_4$. Denote by $A'$ the kernel of the surjection from $V_6$ to $A$ and $\mathbb{P}(A ^{\vee})\subset {\mathbb{P}_5}$ by $\pi_A$. The restriction of $\Omega^1_{{\mathbb{P}_5}}(1)$ to $\pi_A$ is $A'\otimes {\cal O}_{\pi_A} \oplus \Omega^1_{\pi_A}(1)$. From the isotropicity of $\pi_A$ with respect to all the elements of $B$ we see that the restriction of $\alpha$ to $\pi_A$ is the direct sum of the following maps: \begin{center} $ B \otimes {\cal O}_{\pi_A}(-1) \to A' \otimes{\cal O}_{\pi_A}$ and $ \frac{W_4 ^{\vee}}{B} \otimes {\cal O}_{\pi_A}(-1) \to \Omega^1_{\pi_A}(1)$. \end{center} The determinant of the first one gives a cubic curve in $\pi_A\cap X$, and the second map vanishes on a single (residual) point $\mu$ of $\pi_A\cap X$. So we have constructed a morphism from $Z_4$ to $X$: $(A ^{\vee},B) \mapsto \mu$. Moreover, this vanishing shows by specialization of sequence (\ref{resEX}) at the point $\mu$ that the fiber of $E_X ^{\vee}$ at $\mu$ is $A ^{\vee}$. So $Z_4$ and $X$ have the same image in $G(3,V_6 ^{\vee})$, and the proof of the proposition is reduced to the proof of the embedding of $Z_4$ to $G(3,V_6 ^{\vee})$. But the fiber of this morphism over the point of $G(3,V_6 ^{\vee})$ corresponding to $A ^{\vee}$ is a single point because $A ^{\vee}$ is not isotropic for all the elements of $W_4 ^{\vee}$. So this projection of $Z_4$ is one to one, and it must be an isomorphism because the fibers are given by linear conditions. \nolinebreak $\Box$ \subsection{Anticanonical properties} Although it is classical that the canonical class $K_X$ of a Palatini threefold satisfies $K_X^3=-2$ (cf. for instance \cite{Ok}) the following identification and next proposition seem new. \begin{lemma}\label{W4W4v} The canonical line bundle of $X$: $\omega_X$ is isomorphic to ${\cal O}_X(s-2h)$. With the notations \ref{notapala}, we have from the equality $W_4= H^0({\cal O}_S(1))$ a canonical isomorphism: $$H^0(\omega_X ^{\vee}) = W_4 ^{\vee} $$ \end{lemma} {\noindent\it Proof: } The isomorphism $\omega_X\simeq{\cal O}_X(s-2h)$ can be computed directly from the definition \ref{palatini}. We obtain the isomorphism $H^0(\omega_X ^{\vee}) = W_4 ^{\vee} $ from the isomorphism between $X$ and $Z_4$ found in the proposition \ref{Z4isoX} and the fact that $\omega_{Z_4} ^{\vee}$ is the pull back of ${\cal O}_{\mathop{\pi_3 ^\vee}}(1)$. \nolinebreak $\Box$ \begin{proposition}\label{propantican} The linear system $|\omega_X ^{\vee}|$ has no base points and gives a morphism of degree $2$: $$ \begin{CD} X @>2:1>> \mathop{\pi_3 ^\vee} \subset G(3,\bigwedge\limits^2 V_6 )\end{CD} $$ The anticanonical linear system of $X$ contracts $5$ rational curves. These curves have degree $3$ for the embeding of $X$ in ${\mathbb{P}_5}$ and also for the embeding of $X$ in $G(3,V_6 ^{\vee})$. \end{proposition} {\noindent\it Proof: } The contracted curves of this morphism correspond to the case c) of lemma \ref{3casplans}: the planes of a Segre. So they are smooth rational cubic curves in $G(3,V_6 ^{\vee})$. By definition, on such a curve, de divisors $2h$ and $s$ are equivalent because $\omega_X ^{\vee}={\cal O}_X(2h-s)$. So those curves have the same degree with respect to $h$ than to $3h-s$. So the proof will end after the following: \def\bar{F}{\bar{F}} \begin{lemma} Let $\bar{F}$ be the normalized bundle $F(-1)$. The vector space $$H=H^1\left((S^2\bar{F})(-1)\right)$$ has dimension $5$ and it is the kernel of the following map given by the pfaffians of size $4$ of $M$: \begin{equation} \label{eq:kerpfaf44} \begin{CD} 0 @>>> H @>>> \bigwedge\limits^2 V_6 = \bigwedge\limits^4 V_6 ^{\vee} @>>> S^2 W_4 @>>>0 \end{CD} \end{equation} Moreover the ideal of the exceptional locus in $\mathop{\pi_3 ^\vee}$ of the projection $X=Z_4 \to \mathop{\pi_3 ^\vee}$ is given by the $4\times 4$ pfaffians of a skew-symmetric map: $$\begin{CD} H \otimes {\cal O}_{\mathop{\pi_3 ^\vee}} (-1) @>>> H ^{\vee} \otimes {\cal O}_{\mathop{\pi_3 ^\vee}} \end{CD}. $$ \end{lemma} {\noindent\it Proof: } Let $i$ be an isomorphism: $\wedge^2\bar{F} \to {\cal O}_S$. The restriction of $F$ to a plane $P$ is of type c) in lemma \ref{3casplans} if and only if we have $h^1(S^2(\bar{F}_P))=3$. To globalize this condition, let's consider the complex: $$C^\bullet: \begin{CD} 0 @>>> V_6 ^{\vee}\otimes{\cal O}_{\mathop{\pi_3}}(-2) @>M>> V_6\otimes{\cal O}_{\mathop{\pi_3}}(-1) @>>> 0 \end{CD}. $$ It is exact in degree $-1$ with cohomology $\bar{F}$ in degree $0$. The exterior power of $C^\bullet$ tensorized by ${\cal O}_{\mathop{\pi_3}}(2)$ is: $$\begin{CD} 0 @>>> S^2 V_6 ^{\vee}\otimes{\cal O}_{\mathop{\pi_3}}(-2) @>>> V_6 ^{\vee}\otimes V_6\otimes{\cal O}_{\mathop{\pi_3}}(-1) @>>> \bigwedge\limits^2V_6\otimes{\cal O}_{\mathop{\pi_3}} @>>> 0 \end{CD} $$ with cohomology in degree $(-2,-1,0)$: $(0,S^2(\bar{F})(-1),(\wedge^2(\bar{F})(2)))$. So the hyper-cohomology's spectral sequence of this complex gives the exact sequence (\ref{eq:kerpfaf44}), the dimension of $H$, and the vanishings $$h^0(S^2(\bar{F})(-1))=h^2(S^2(\bar{F})(-1))=h^0(S^2(\bar{F}))=h^2(S^2(\bar{F}))=0.$$ Now consider the point/plane incidence variety $I_3\subset \mathop{\pi_3 ^\vee}\times \mathop{\pi_3}$ and denote by $p ^{\vee}_{3}$ and $p_3$ the first and second projections of this product. We have the exact sequence: $$ \begin{CD} 0@>>> {\cal O}_{\mathop{\pi_3 ^\vee}}(-1)\boxtimes S^2\bar{F}(-1) @>>> {\cal O}_{\mathop{\pi_3 ^\vee}}\boxtimes S^2\bar{F} @>>> p_3^*(S^2\bar{F}) @>>> 0 \end{CD} $$ From the Leray's spectral sequence and the above vanishings, we have the exact sequence: {\small\minCDarrowwidth1em $$ \begin{CD} 0@>>> p ^{\vee}_{3*}(p_3^*(S^2\bar{F})) @>>> H^1(S^2(\bar{F})(-1))\otimes {\cal O}_{\mathop{\pi_3 ^\vee}}(-1)@>d_M>> H^1(S^2(\bar{F}))\otimes {\cal O}_{\mathop{\pi_3 ^\vee}} @>>> R^1p ^{\vee}_{3*}(p_3^*(S^2\bar{F})) @>>> 0 \end{CD} $$ } Let's now explain how to consider the map $d_M$ as a skew-symmetric map. The isomorphism $i$ gives a symmetric isomorphism $i': S_2(\bar{F}) \to S_2(\bar{F} ^{\vee})$ so the following square is commutative: $$ \begin{CD} (S_2\bar{F})(-1) \otimes S_2\bar{F} @>i'\otimes id>> (S_2\bar{F} ^{\vee})(-1) \otimes S_2\bar{F}\\ @Vid \otimes i'VV @VV{\tau}V \\\ (S_2\bar{F})(-1) \otimes S_2\bar{F} ^{\vee} @>{\tau'}>> {\cal O}_S(-1) \end{CD} $$ The cup-product $H^1\left((S_2\bar{F})(-1)\right)\otimes H^1\left((S_2\bar{F})(-1)\right) \to H^2\left((S_2\bar{F}\otimes S_2\bar{F})(-2)\right)$ is anti-commutative, so for any $z\in W_4$ the following square is also anti-commutative: $$ \begin{CD} H^1\left((S_2\bar{F})(-1)\right) \otimes H^1\left((S_2\bar{F})(-1)\right) @>d_{M,z}\otimes id>> H^1\left(S_2\bar{F}\right) \otimes H^1\left((S_2\bar{F})(-1)\right)\\ @Vid \otimes d_{M,z}VV @VV{\cup}V \\ H^1\left((S_2\bar{F})(-1)\right) \otimes H^1\left(S_2\bar{F}\right) @. H^2\left(S_2\bar{F}\otimes (S_2\bar{F})(-1)\right) \\ @V{\cup}VV @VV{\overline{\tau\circ(i'\otimes id)}}V \\ H^2\left((S_2\bar{F})(-1) \otimes S_2\bar{F}\right) @>\overline{\tau'\circ(id\otimes i')}>> H^2({\cal O}_S(-1)) \end{CD} $$ In conclusion, the composition: $$ \begin{CD} H \otimes {\cal O}_{\mathop{\pi_3 ^\vee}}(-1)@>d_M>>H^1(S_2\bar{F})\otimes {\cal O}_{\mathop{\pi_3 ^\vee}}@>\bar{i'}>>H^1(S_2(\bar{F} ^{\vee})\otimes {\cal O}_{\mathop{\pi_3 ^\vee}})@>Serre's\ duality>> H ^{\vee}\otimes {\cal O}_{\mathop{\pi_3 ^\vee}} \end{CD} $$ is skew-symmetric and the lemma is proved. Indeed, the type c) cases correspond to the locus where this map has rank at most $2$. \nolinebreak $\Box$ \begin{definition}\label{defPHI1} Let $\Sigma_5$ be the symmetric product of order $5$ of $\mathop{\pi_3 ^\vee}$. We define the rational map $\Phi_1$ to be: $$ \Phi_1: \begin{array}[t]{ccc} (W_4\otimes \bigwedge\limits^2 V_6)^{sm}/GL(V_6) & \dashrightarrow & \mathbb{P}(S^3(W_4))\times \Sigma_5\\ \alpha & \mapsto & (S,(h_0\dots h_4)) \end{array} $$ where $S$ is the pfaffian cubic surface defined by $\alpha$, and $h_0,\dots,h_4$ are the five linear sections of $S$ defined in proposition \ref{propantican}. \end{definition} In section \ref{W2V6} we will understand the image of this map. \subsection{Palatini threefolds and endomorphisms} Although this part is not required by the main theorem, let's briefly describe here some connected remarks. The exceptional geometric properties of a Palatini threefold are classically considered as natural generalizations of what happens to a Veronese $\mathcal{V}$ surface embeded in $\mathbb{P}_4$. Note for instance, in the Veronese situation, the sequence \ref{resopalaP5} is replaced by: $$ \begin{CD} 0@>>> W_3 ^{\vee} \otimes {\cal O}_{\mathbb{P}_4} @>\alpha>> \Omega^1_{\mathbb{P}_4}(2h) @>>> I_{\mathcal{V}}(3h) @>>> 0 \end{CD}. $$ But the main difference is that in the theory of Severi varieties the embedding of $\mathcal{V}$ by the complete linear system $|{\cal O}_{\mathcal{V}}(h)|$ is understood from an interpretation in terms of matrices of size $3\times 3$ of rank $1$. For a Palatini threefold, there is no similar result to describe the embedding by the complete linear system $|{\cal O}_X(2h)|$. The following remark could be a first step in this direction: \begin{remark}\label{endopala} The restriction of the line bundle $\omega_X ^{\vee} \boxtimes {\cal O}_X(s)$ to the diagonal of $X\times X$ gives the natural inclusions: $$W_4 ^{\vee} \otimes W_4 \subset H^0({\cal O}_X(2h))$$ In other words, the embedding of a Palatini threefold $X$ with $|{\cal O}_X(2h)|$ has a canonical projection in $\mathbb{P}(W_4 ^{\vee} \otimes W_4)$, and the image of $X$ by this projection is included in the endomorphisms of $W_4$ of rank $1$. \end{remark} {\noindent\it Proof: } It's straighforward from lemma \ref{W4W4v}. \nolinebreak $\Box$ \section{Geometry in $\bigwedge\limits^2 V_6$}\label{W2V6} \subsection{Projections from linear spaces} The Grassmannian variety $G(2,6)$ is one of the $4$ Severi varieties. It is well known to have the exceptional property that its projection from a general line has a unique triple point (cf \cite{I-M}, \cite{Z}). Here, we prove that it has the same property with projection from general $\mathbb{P}_3$ and points of multiplicity $5$: \begin{proposition}\label{P4-5sec} Denote by $\mathcal{U}_5$ the subspace of $G(5,\bigwedge\limits^2 V_6)$ defined by the five dimensional vector spaces such that the intersection of their projectivisation with $G(2,V_6)$ is $5$ linearly independent distinct points. Let $W_4 ^{\vee}$ be a general $4$-dimensional subspace of $\bigwedge\limits^2 V_6$, then there is a unique element of $\mathcal{U}_5$ containing $W_4 ^{\vee}$. \end{proposition} {\noindent\it Proof: } First remark that the incidence variety $$I_{4,5}=\{(W_4 ^{\vee},W_5 ^{\vee})| W_4 ^{\vee}\subset W_5 ^{\vee} \subset \bigwedge\limits^2 V_6, W_5 ^{\vee} \in \mathcal{U}_5\}$$ has the same dimension as $G(4,\bigwedge\limits^2V_6)$, so we have to prove that the natural projection is birational. So, consider a general element $W_4 ^{\vee}$ in the image of this projection, and chose an element $W_5 ^{\vee}$ such that $(W_4 ^{\vee},W_5 ^{\vee})\in I_{4,5}$. Denote by $\mathop{\pi_3},\pi_4$ their projectivisation. The vector space $H^0(I_{\mathop{\pi_3} \cup G(2,V_6)}(2))$ is the kernel of the map $\bigwedge\limits^4 V_6 ^{\vee}\to S_2 W_4$. So it has dimension $5$. Now remark that we also have $h^0(I_{\pi_4 \cup G(2,V_6)}(2))=5$ because the ideal of the $5$ points $\pi_4 \cap G(2,V_6)$ in $\pi_4$ is a $10$ dimensional space of quadrics. So we proved that $\pi_4$ must be in all the quadrics of $H^0(I_{\mathop{\pi_3} \cup G(2,V_6)}(2))$. It gives the following linear conditions satisfied by any $W_5 ^{\vee}$ of $\mathcal{U}_5$ containing $W_4 ^{\vee}$: $$ W_5 ^{\vee} \subset \bigcap\limits_{q \in H^0(I_{\mathop{\pi_3} \cup G(2,V_6)}(2)) } (W_4 ^{\vee})^{\bot_q}$$ where $\bot_q$ denotes the orthogonal with respect to the quadratic form $q$ on $\bigwedge^2 V_6$. So unicity of $W_5 ^{\vee}$ will be a corollary of existence of an exemple of $W_4$ such that $\bigcap\limits_{q \in H^0(I_{\mathop{\pi_3} \cup G(2,V_6)}(2)) } (W_4 ^{\vee})^{\bot_q}$ has dimension $5$ as it is the case in the following: \begin{example} Let's consider a basis $(\eps_i)$ of $V_6$, and the $5$ elements $$u_0=\eps_0\wedge \eps_3, u_1=\eps_1 \wedge \eps_4, u_2=\eps_2 \wedge \eps_5, u_3=(\eps_0+\eps_1+\eps_2)\wedge (\eps_4+\eps_3+\eps_5),u_4=(\eps_1+\eps_4+\eps_2)\wedge (\eps_3+\eps_1+\eps_5).$$ Denote by $W_5 ^{\vee}$ the $5$ dimensional vector space spanned by the $(u_i)$ and $$W_4 ^{\vee}=\{\sum_{0\leq i \leq 4} \lambda_i.u_i | \sum_{0\leq i \leq 4} \lambda_i =0\}.$$ Then $\bigcap\limits_{q \in H^0(I_{\mathop{\pi_3} \cup G(2,V_6)}(2)) } (W_4 ^{\vee})^{\bot_q}$ has dimension $5$. \end{example} {\noindent\it Proof: } We can compute with \cite{Macaulay2} that $H^0(I_{\mathop{\pi_3}\cup G(2,V_6)}(2))$ is generated by the following five quadrics in Plucker coordinates: \begin{itemize} \item[.] ${p}_{({3},{4})} {p}_{(1,{5})}-{p}_{(1,{4})} {p}_{({3},{5})}+{p}_{(1,{3})} {p}_{({4},{5})}$ \item[.]$ {p}_{(1,{2})} {p}_{(0,{5})}-{p}_{({2},{4})} {p}_{(0,{5})}-{p}_{(0,{2})} {p}_{(1,{5})}+{p}_{({2},{3})} {p}_{(1,{5})}+{p}_{(0,1)} {p}_{({2},{5})}-{p}_{(1,{3})} {p}_{({2},{5})}+{p}_{(0,{4})} {p}_{({2},{5})}-{p}_{({3},{4})} {p}_{({2},{5})}+{p}_{(1,{2})} {p}_{({3},{5})}+{p}_{({2},{4})} {p}_{({3},{5})}-{p}_{(0,{2})} {p}_{({4},{5})}-{p}_{({2},{3})} {p}_{({4},{5})}$ \item[.] $ {p}_{({2},{3})} {p}_{(0,{4})}-{p}_{(0,{3})} {p}_{({2},{4})}+{p}_{(0,{2})} {p}_{({3},{4})}-{p}_{(1,{3})} {p}_{(0,{5})}+{p}_{({2},{4})} {p}_{(0,{5})}-{p}_{({3},{4})} {p}_{(0,{5})}+{p}_{(0,{3})} {p}_{(1,{5})}-{p}_{({2},{3})} {p}_{(1,{5})}+{p}_{(1,{3})} {p}_{({2},{5})}-{p}_{(0,{4})} {p}_{({2},{5})}+{p}_{({3},{4})} {p}_{({2},{5})}-{p}_{(0,1)} {p}_{({3},{5})}-{p}_{(1,{2})} {p}_{({3},{5})}+{p}_{(0,{4})} {p}_{({3},{5})}-{p}_{({2},{4})} {p}_{({3},{5})} +{p}_{(0,{2})} {p}_{({4},{5})}-{p}_{(0,{3})} {p}_{({4},{5})}+{p}_{({2},{3})} {p}_{({4},{5})}$ \item[.] $ {p}_{(1,{2})} {p}_{(0,{4})}-{p}_{(0,{2})} {p}_{(1,{4})}+{p}_{(0,1)} {p}_{({2},{4})}-{p}_{({2},{4})} {p}_{(0,{5})}+{p}_{({2},{3})} {p}_{(1,{5})}-{p}_{(1,{3})} {p}_{({2},{5})}+{p}_{(0,{4})} {p}_{({2},{5})}-{p}_{({3},{4})} {p}_{({2},{5})}+{p}_{(1,{2})} {p}_{({3},{5})}+{p}_{({2},{4})} {p}_{({3},{5})}-{p}_{(0,{2})} {p}_{({4},{5})}-{p}_{({2},{3})} {p}_{({4},{5})}$ \item[.] $ {p}_{(1,{2})} {p}_{(0,{3})}-{p}_{(0,{2})} {p}_{(1,{3})}+{p}_{(0,1)} {p}_{({2},{3})}-{p}_{({2},{4})} {p}_{(0,{5})}+{p}_{({3},{4})} {p}_{(0,{5})}+{p}_{({2},{3})} {p}_{(1,{5})}-{p}_{(1,{3})} {p}_{({2},{5})}+{p}_{(0,{4})} {p}_{({2},{5})}-{p}_{({3},{4})} {p}_{({2},{5})}+{p}_{(1,{2})} {p}_{({3},{5})}-{p}_{(0,{4})} {p}_{({3},{5})}+{p}_{({2},{4})} {p}_{({3},{5})}-{p}_{(0,{2})} {p}_{({4},{5})}+{p}_{(0,{3})} {p}_{({4},{5})}-{p}_{({2},{3})} {p}_{({4},{5})}$ \end{itemize} and check that the ideal of the orthogonal of $\mathop{\pi_3}$ with respect to these $5$ quadrics is generated by the $10$ independant equations: $ ({p}_{({3},{5})},{p}_{(0,{5})}-{p}_{(1,{5})}+{p}_{({4},{5})},{p}_{({3},{ 4})}+{p}_{({4},{5})},{p}_{({2},{4})}-{p}_{(1,{5})}+{p}_{({4},{5})},{p }_{(0,{4})}-{p}_{(1,{5})}+{p}_{({4},{5})},{p}_{({2},{3})}-{p}_{(1,{5 })},{p}_{(1,{3})}-{p}_{(1,{5})},{p}_{(1,{2})}+{p}_{({4},{5})},{p}_{(0,{2 })},{p}_{(0,1)})$. So this example completes the proof of the birationality of the projection from $I_{4,5}$ to $G(4,\bigwedge\limits^2 V_6)$. So we have proved proposition \ref{P4-5sec}. \nolinebreak $\Box$ \begin{corollary}\label{defPHI2} With notations of definition \ref{inscribed}, we can define the rational map $\Phi_2$ by: $$ \Phi_2: \begin{array}[t]{ccc} (W_4\otimes \bigwedge\limits^2 V_6)^{sm}/GL(V_6) & \dashrightarrow & \mathcal{H}\\ \alpha & \mapsto & (S,(H_0\dots H_4)) \end{array} $$ where $S$ is the pfaffian cubic surface defined by $\alpha$, and $H_i$ is defined like this: From proposition \ref{P4-5sec}, consider the five points $(u_i)_{0\leq i \leq 4}$ of $G(2,V_6)$ such that $\mathop{\pi_3}$ is in the linear span $<(u_i)_{0\leq i \leq 4}>$. Then take: $$H_i=\mathop{\pi_3}\cap <(u_j)_{0\leq j\leq 4, j\neq i} >$$ \end{corollary} {\noindent\it Proof: } After proposition \ref{P4-5sec}, we only have to explain why $H_0,\dots,H_4$ is inscribed on $S$. But for $\{i_0,\dots,i_4\}=\{0,\dots,4\}$ the point $H_{i_0}\cap H_{i_1}\cap H_{i_2}$ is on the line $(u_{i_3},u_{i_4})$ so it corresponds to a matrix of rank $4$ and is on $S$. \nolinebreak $\Box$ \begin{remark}\label{Hrational} The variety $\mathcal{H}$ is rational of dimension $24$. \end{remark} {\noindent\it Proof: } Let $\Sigma_5'$ be the image of $\mathcal{H}$ in $|{\cal O}_{\mathop{\pi_3}}(5)|$ by the second projection. It is an openset of the symmetric product $\Sigma_5$ defined in \ref{defPHI1}. So it is a rational $15$-dimensional variety (cf \cite{GKZ} Th 2.8 p 137). The partial derivatives of order $2$ of any element of $\Sigma_5'$ are linearly independent cubic forms. So they give a rank $10$ subsheaf $\mathcal{F}_2$ of $H^0({\cal O}_{\mathop{\pi_3}}(3))\otimes {\cal O}_{\Sigma_5'}$ locally free with respect to Zariski's topology. Now remark that $\mathcal{H}$ is the openset of $\mathbb{P}(\mathcal{F}_2)$ corresponding to smooth cubic surfaces. So $\mathcal{H}$ is rational of dimension $24$. \nolinebreak $\Box$ \subsection{An explicit formula and proof of theorem \ref{mainbir}}\label{sectionexplicit} Surprisingly, we are able to give in this section an explicit formula. Recently, a explicit result was also found by F. Tanturri in \cite{T}: An algorithm to obtain a pfaffian representation from a cubic equation. The two main difference, are the following: -first he wants to find any pfaffian representation of $S$, but here we need to find a unique point in the moduli space. -The construction starts with five points on $S$, so it is a problem of extending the $5$ by $5$ skew-symmetric matrix of the resolution of the $5$ points to a $6$ by $6$ one with pfaffian $S$, while we start with an inscribed pentahedron. \begin{lemma}\label{equationnorm} Let $(x_i)_{0\leq i \leq 3}$ be a basis of $W_4$, and $\mathcal{A}_9$ be the following subspace of $\mathbb{C}^{10}\times \mathbb{P}_4$: $$ \mathcal{A}_9=\left\{\left((a_{i,j,k})_{0\leq i<j<k\leq 4},(b_i)_{0\leq i\leq 4} \right) \left| \begin{array}[c]{l} a_{0,1,4}=1\mbox{ and for } 0\leq i \leq 4,\ b_i\neq 0, \\ \mbox{and for } 0\leq i<j<k\leq 3,\ a_{i,j,k}=1 \end{array}\right. \right\}.$$ Then the following map is birational: \begin{equation} \begin{array}[c]{ccc} PGL_4\times \mathcal{A}_9 & \to & \mathcal{H}_{ord}\\ (P,\left((a_{i,j,k})_{0\leq i<j<k\leq 4},(b_i)_{0\leq i\leq 4} \right)) & \mapsto & (S,(H_0,\dots,H_4)) \end{array} \end{equation} where $$\sum_{0\leq i<j<k\leq 4} a_{i,j,k}.w_i.w_j.w_k =0, $$ is an equation of $S$, and for all $0\leq i \leq 4$, $w_i=0$ is an equation of $H_i$ with the following equalities: $w_4=\sum\limits_{i=0}^3 \frac{b_4.w_i}{b_i}$, {\footnotesize$\left(\begin{array}[c]{l} w_0\\ w_1\\ w_2\\ w_3 \end{array}\right)=P. \left(\begin{array}[c]{l} x_0\\ x_1\\ x_2\\ x_3 \end{array}\right) $}. \end{lemma} {\noindent\it Proof: } Let $\Pi=(H_0,\dots,H_4)$ be an ordered pentahedron and $P'$ be the unique projective transformation that sends the ordered pentahedron $(x_0,x_1,x_2,x_3,x_0+x_1+x_2+x_3)$ to $(H_0,\dots,H_4)$. Denote by $h_i$ the equation of $H_i$ defined by: {\footnotesize$\left(\begin{array}[c]{l} h_0\\ h_1\\ h_2\\ h_3 \end{array}\right)=P'. \left(\begin{array}[c]{l} x_0\\ x_1\\ x_2\\ x_3 \end{array}\right) $} and $h_4=\sum_{i=0}^3 h_i$. Cubic surfaces $S$ such that $(S,\Pi)$ is in $\mathcal{H}_{ord}$ are the smooth surfaces with equation: $$ \sum_{0\leq i<j<k\leq 4} A_{i,j,k}.h_i.h_j.h_k =0,\ (A_{i,j,k})_{(0\leq i < j < k \leq 4)} \in \mathbb{P}_9.$$ Now remark that the map: $$ \begin{array}[c]{lll} \mathcal{A}_9 & \to \mathbb{P}_9\\ (a,b) & \mapsto & (A_{i,j,k}=a_{i,j,k}.b_i.b_j.b_k)_{0\leq i <j <k \leq 4} \end{array} $$ is birational because we can compute its inverse with the following formulas\footnote{If one works with affine spaces instead of $\mathbb{P}_9$ and $\mathbb{P}_4$, then one needs to extract a cubic root to solve the equalities.}: $$ \frac{b_0}{b_3}=\frac{A_{0,1,2}}{A_{1,2,3}},\frac{b_1}{b_3}=\frac{A_{0,1,2}}{A_{0,2,3}},\frac{b_2}{b_3}=\frac{A_{0,1,2}}{A_{0,1,3}},\frac{b_4}{b_3}=\frac{A_{0,1,4}}{A_{0,1,3}}, a_{i,j,4}=\frac{A_{i,j,4}}{A_{i,j,3}}\cdot \frac{A_{0,1,3}}{A_{0,1,4}}.$$ So we obtain the lemma from the equalities: $0\leq i \leq 4, w_i=h_i.b_i$ with $P$ defined by the product of the diagonal matrix $(\frac{b_0}{b_4},\dots,\frac{b_3}{b_4})$ with $P'$. \nolinebreak $\Box$ \begin{definition}\label{defexplicit} Let $\mathcal{A}'_9$ be the set of triples $(a,b,u)$ such that $(a,b)$ is an element of $\mathcal{A}_9$ defining a smooth cubic surface: $$\sum_{0\leq i<j<k\leq 4} a_{i,j,k}.w_i.w_j.w_k =0,\ w_4=\sum\limits_{i=0}^3 \frac{b_4.w_i}{b_i},$$ and $u$ is a root of the following equation in $X$: $$X^{2}+X\cdot (1+a_{0,2,4}-a_{0,3,4})+a_{0,2,4}=0.$$ Denote by $v=-(1+a_{0,2,4}-a_{0,3,4})-u$ the other one and define: $$e_1=a_{0,2,4}+a_{1,2,4}-a_{2,3,4},\ e_2=1+a_{1,2,4}-a_{1,3,4},\ e_3=(-a_{1,2,4}+a_{1,3,4}-1) v-a_{1,2,4}-a_{0,2,4}+a_{2,3,4} $$ $$M_4=\left(\begin{array}{cccccc} 0 & u & -1 & a_{1,2,4} & e_1 & e_2 \\ -u & 0 & 0 & 0 & a_{0,2,4} & -u \\ 1 & 0 & 0 & 0 & -v & 1 \\ -a_{1,2,4} & 0 & 0 & 0 & a_{1,2,4} v & -a_{1,2,4} \\ -e_1 & -a_{0,2,4} & v & -a_{1,2,4} v & 0 & e_3 \\ -e_2 & u & -1 & a_{1,2,4} & -e_3 & 0 \end{array}\right) $$ $$M_{0123}=\left(\begin{array}{cccccc} 0 & 0 & 0 & w_{0}+w_{3} & w_{3} & w_{3} \\ 0 & 0 & 0 & w_{3} & w_{1}+w_{3} & w_{3} \\ 0 & 0 & 0 & w_{3} & w_{3} & w_{2}+w_{3} \\ -w_{0}-w_{3} & -w_{3} & -w_{3} & 0 & 0 & 0 \\ -w_{3} & -w_{1}-w_{3} & -w_{3} & 0 & 0 & 0 \\ -w_{3} & -w_{3} & -w_{2}-w_{3} & 0 & 0 & 0 \end{array}\right) $$ \end{definition} \begin{theorem}\label{explicit} For a generic element $(P,(a,b,u))$ of $PGL_4\times \mathcal{A}'_9$, the element $M$ of $(W_4\otimes \bigwedge\limits ^2 V_6)$ defined by $M=M_{0123}+w_4 M_4$ is such that: $\Phi_1(M)=\Phi_2(M)=(S,\Pi)$ where the equation of $S$ and $\Pi$ are given by the formula in lemma \ref{equationnorm}. \end{theorem} {\noindent\it Proof: } The difficulty was to find $M_4$. It was done by tracking the rational cubic curve in ${\mathbb{P}_5}$ associated to the plane $w_4=0$ in proposition \ref{propantican}. But now that we have found $M_4$, it is much easier to check that $M$ safisties the required properties. NB: To obtain a more compact presentation, we have glued the indexes of the $a_{i,j,k}$ in the next formulas. - First, one can check that the pfaffian of $M$ is $$ a_{024} w_{0} w_{2} w_{4}+ a_{034} w_{0} w_{3} w_{4}+a_{234} w_{2} w_{3} w_{4} +a_{124} w_{1} w_{2} w_{4} +a_{134} w_{1} w_{3} w_{4} + w_{0} w_{1} w_{4}+\sum_{0\leq i<j<k\leq 3} w_i w_j w_k$$ - Now to prove that $\Phi_2(\alpha)=(S,\Pi)$ we just have to remark that $M_4$ has rank $2$, and also the $4$ values of $M_{0123}$ at the points where $(w_0,w_1,w_2,w_3)$ take the values $(1,0,0,0)$, $(0,1,0,0)$, $(0,0,1,0)$, $(0,0,0,1)$. - To obtain that $\Phi_1(\alpha)=(S,\Pi)$ we need to find $5$ elements $(P_i)$ of $GL(V_6)$ such that $\t{P_i}.M.P_i = \left(\begin{array}[c]{cc} 0 & A_i\\ -A_i & 0 \end{array}\right) $ where $A_i$ are $3$ by $3$ symmetric matrices with linear entries. We found the following ones easily, $$P_4=Id, P_3=\left(\begin{array}{cccccc} 0 & 0 & 0 & 0 & 0 & 1 \\ \frac{v}{u} & 0 & 0 & \frac{(-a_{024}-a_{124}+a_{234})}{u} & 0 & 0 \\ 0 & 1 & 0 & 0 & 1+a_{124}-a_{134} & 0 \\ 0 & 0 & \frac{-1}{a_{124}} & 0 & 0 & 0 \\ 0 & 0 & 0 & 1 & 0 & 0 \\ 0 & 0 & 0 & 0 & 1 & 0 \end{array}\right) $$ but the next ones only after understanding that we should use the $SL_2\times SL_2 \times SL_2$ action that preserves the $3$ marked lines in the intersection of the two Segre $\mathbb{P}_1\times \mathbb{P}_2$ defined by $w_i=0$ and $w_4=0$ $$P_1=\left(\begin{array}{cccccc} 0 & 0 & \frac{-a_{024}}{a_{234}} & 0 & 0 & 0 \\ \frac{(-u) (a_{024}+u)}{a_{024} (u+1)} & 1 & \frac{a_{024}}{a_{234}} & \frac{a_{024} a_{134}-a_{024} u-a_{024}+u a_{234}}{a_{024} (u+1)} & \frac{-a_{124}}{u} & 0 \\ \frac{u (a_{024}+u)}{a_{024} (u+1)} & 0 & 0 & \frac{-a_{024} a_{134}+a_{024} u+a_{024}-u a_{234}}{a_{024} (u+1)} & 0 & 0 \\ 0 & 0 & \frac{u}{a_{234}} & 0 & 0 & -1 \\ 0 & 0 & \frac{-u}{a_{234}} & -1 & 1 & 1 \\ 0 & 0 & 0 & 1 & 0 & 0 \end{array}\right) $$ $$P_2= \left(\begin{array}{cccccc} \frac{1}{a_{134}} & 0 & 0 & 0 & 0 & 0 \\ 0 & \frac{1+v}{u+1} & 0 & 0 & \frac{-v-a_{024}+a_{234}+v a_{134}}{u+1} & 0 \\ -\frac{1}{a_{134}} & \frac{-1-v}{u+1} & 1 & 0 & \frac{v+a_{024}-a_{234}-v a_{134}}{u+1} & a_{124} \\ \frac{1}{a_{134}} & 0 & 0 & 1 & 0 & 0 \\ 0 & 0 & 0 & 0 & 1 & 0 \\ -\frac{1}{a_{134}} & 0 & 0 & -1 & -1 & 1 \end{array}\right) $$. $$P_0= \left(\begin{array}{cccccc} a_{124} & \frac{-\left(a_{024} u+a_{024}-u a_{234}\right)^{2}}{u^{2} a_{234}} & -a_{134} & \frac{-a_{124} a_{024}}{a_{024} u+a_{024}-u a_{234}} & 0 & \frac{a_{024} a_{134}-a_{024} u-a_{024}+u a_{234}}{a_{024} u+a_{024}-u a_{234}} \\ 0 & \frac{\left(a_{024} u+a_{024}-u a_{234}\right)^{2}}{u^{2} a_{234}} & 0 & 0 & 0 & 0 \\ 0 & 0 & a_{134} & 0 & 0 & \frac{-a_{024} a_{134}+a_{024} u+a_{024}-u a_{234}}{a_{024} u+a_{024}-u a_{234}} \\ 1 & \frac{(u+1) (a_{024} u+a_{024}-u a_{234})}{a_{234} u} & 0 & \frac{u (a_{024}-a_{234})}{a_{024} u+a_{024}-u a_{234}} & -1 & -1 \\ 0 & \frac{(-u-1) (a_{024} u+a_{024}-u a_{234})}{a_{234} u} & 0 & 0 & 1 & 0 \\ 0 & 0 & 0 & 0 & 0 & 1 \end{array}\right) $$ and we have proved theorem \ref{explicit}. \nolinebreak $\Box$ \bigskip We are now able to obtain a more explicit version of theorem \ref{mainbir} stated in introduction. \begin{corollary}\label{bir} Maps $\Phi_1$ and $\Phi_2$ coincide on a open set, and give birational maps: $$(W_4\otimes \bigwedge\limits^2 V_6)^{sm}/GL(V_6) \dashrightarrow \mathcal{H}.$$ \end{corollary} {\noindent\it Proof: } First remark that both spaces are irreducible of dimension $24$. Now consider with notations of definition \ref{defexplicit} the following map: \begin{center} $ \begin{array}[c]{lll} PGL_4\times \mathcal{A}'_9 & \to & W_4\otimes \bigwedge\limits^2 V_6\\ (P,a,b,u) & \mapsto & M_{0123}+w_4.M_4 \end{array} $, where $w_4=\sum\limits_{i=0}^3 \frac{b_4.w_i}{b_i}$, and {\footnotesize$\left(\begin{array}[c]{l} w_0\\ w_1\\ w_2\\ w_3 \end{array}\right)=P. \left(\begin{array}[c]{l} x_0\\ x_1\\ x_2\\ x_3 \end{array}\right) $}. \end{center} and denote by $f$ its composition with the canonical projection from $ (W_4\otimes \bigwedge\limits^2 V_6)^{sm}$ to $ (W_4\otimes \bigwedge\limits^2 V_6)^{sm}/GL(V_6)$. The map $PGL_4\times\mathcal{A}_9' \to PGL_4\times \mathcal{A}_9$ has degree $2$ because of the permutation of $u$ and $v$, and the rational map $PGL_4\times \mathcal{A}_9$ to $\mathcal{H}$ has degree $5!$ from the choice of the order and lemma \ref{equationnorm}. So we have from theorem \ref{explicit} the commutative diagram of rational maps: $$ \begin{CD} PGL_4\times\mathcal{A}_9' @>2:1>> PGL_4\times \mathcal{A}_9 @>1:1>> \mathcal{H}_{ord} \\ @VVfV @. @VV(5!):1V\\ (W_4\otimes \bigwedge\limits^2 V_6)^{sm}/GL(V_6) @.@>\Phi_1>\Phi_2> \mathcal{H} \end{CD} $$ So $\Phi_1$ and $\Phi_2$ are dominant and coincide on an open set, and we just have to prove that $f$ has degree $2.(5!)$ also. We will do this by providing an example of $(S,\Pi)\in \mathcal{H}$ such that the permutation of $u$ and $v$, and the permutations of the elements of $\Pi$ can be obtained by the action of $GL(V_6)$. It is more convenient to take an example where all the elements in the preimage of $(S,\Pi)$ in $PGL_4\times \mathcal{A}'_9$ have all the same values for $(a)$ and $(b)$. So we end the proof with the following invariant example: \begin{example}(Klein-Sylvester) With the following values: $u=e^{\frac{2i\pi}{3}}, v=e^{\frac{-2i\pi}{3}}.$ for $0\leq i < j < k \leq 4,\ a_{i,j,k}=1$. The permutation of $u$ with $v$, and also the permutations of the $(w_i)_{0\leq i\leq 4}$ can be obtained from the action of $GL(V_6)$ on $M=M_{0123}+w_4.M_4$. Note that if we add the conditions $b_i=-b_4$ for $0 \leq i \leq 3$, this is the case of the Klein cubic with its Sylvester Pentahedron). \end{example} {\noindent\it Proof: } Denote by $P_T=I_3\otimes \left(\begin{array}[c]{cc} t_0 & t_1\\ t_2 & t_3 \end{array}\right) $ the matrix {\footnotesize$\left(\begin{array}{cccccc} t_{0} & 0 & 0 & t_{1} & 0 & 0 \\ 0 & t_{0} & 0 & 0 & t_{1} & 0 \\ 0 & 0 & t_{0} & 0 & 0 & t_{1} \\ t_{2} & 0 & 0 & t_{3} & 0 & 0 \\ 0 & t_{2} & 0 & 0 & t_{3} & 0 \\ 0 & 0 & t_{2} & 0 & 0 & t_{3} \end{array}\right) $} and remark that $\t{P_T}M_{0123} P_T = M_{0123}$ when $\left|\begin{array}[c]{cc} t_0 & t_1\\ t_2 & t_3 \end{array}\right|=1$. For a square matrix $\mathcal{T}$, let $D_{\mathcal{T}}$ be the block diagonal matrix $\left(\begin{array}[c]{cc} \mathcal{T}& 0\\ 0 & \mathcal{T} \end{array}\right)$. So we will first use matrices like $D_\mathcal{T}$ to obtain the desired form in the plane $w_4=0$ and then correct the last matrix with $P_T$. We found the following matrices: \begin{itemize} \item[.] Permutation of $u$ and $v$: $P_{uv}=I_3\otimes\left(\begin{array}{cc} \frac{i.u}{\sqrt{2}} & \frac{\sqrt{6}}{2} \\ \frac{-\sqrt{6}}{2} & \frac{i.v}{\sqrt{2}} \end{array}\right) $ then $\t{P_{uv}} M_4 P_{uv} = \overline{M_4}.$ \item[.] $ T_{01}=\left(\begin{array}{ccc} 0 & 1 & 0 \\ 1 & 0 & 0 \\ 0 & 0 & 1 \end{array}\right) $, and $P_{01}=I_3\otimes \left(\begin{array}{cc} \frac{u}{\sqrt{2}} & \frac{v+2}{\sqrt{2}} \\ \frac{-1}{\sqrt{2}} & \frac{-u}{\sqrt{2}} \end{array}\right)$, the conjugation $\t{(D_{T_{01}}.P_{01})}.M.(D_{T_{01}}.P_{01})$ permutes $w_0$ and $w_1$. \item[.] $T_{02}=\left(\begin{array}{ccc} 0 & 0 & 1 \\ 0 & 1 & 0 \\ 1 & 0 & 0 \end{array}\right) $, $P_{02}=I_3 \otimes \left(\begin{array}{cc} \frac{e^{\frac{i\pi}{3}}}{\sqrt{2}}& \frac{-i \sqrt{6}}{2} \\ \frac{e^{\frac{-i\pi}{3}}}{\sqrt{2}}& \frac{v}{\sqrt{2}} \end{array}\right) $, the conjugation $\t{(D_{T_{02}}.P_{02})}.M.(D_{T_{02}}.P_{02})$ permutes $w_0$ and $w_2$. \item[.] $T_{03}=\left(\begin{array}{ccc} 1 & 1 & 1 \\ 0 & -1 & 0 \\ 0 & 0 & -1 \end{array}\right) $, $P_{03}=I_3 \otimes \left(\begin{array}{cc} \frac{i \sqrt{6}}{2} & \frac{-u}{\sqrt{2}}\\ \frac{-v}{\sqrt{2}} & \frac{-i \sqrt{6}}{2} \end{array}\right) $, the conjugation $\t{(D_{T_{03}}.P_{03})}.M.(D_{T_{03}}.P_{03})$ permutes $w_0$ and $w_3$. \item[.] $T_{34}= \left(\begin{array}{ccc} 0 & -\frac{1}{v} & 0 \\ 0 & 0 & 1 \\ 1 & 0 & 0 \end{array}\right)$ , $P_{34}=I_3\otimes \left(\begin{array}{cc} -\frac{v}{\sqrt{2}}& -\frac{i \sqrt{6}}{2}\\ -\frac{i \sqrt{6}}{2} &\frac{u}{\sqrt{2}} \end{array}\right)$, then $\t{(P_3 D_{T_{34}} P_{34})}.M.(P_3 D_{T_{34}} P_{34})$ permutes $w_4$ and $w_3$ with the matrix $P_3$ defined in theorem \ref{explicit}. \end{itemize} This completes the proof because we have provided a generating set of the permutations. \nolinebreak $\Box$ So corollary \ref{bir} is proved and it implies theorem \ref{mainbir} from remark \ref{Hrational}. \nolinebreak $\Box$ \subsection*{A normal form for $5$ general lines in $\mathbb{P}_5$} Explicit forms of definition \ref{defexplicit} and theorem \ref{explicit} have the the following straighforward translation, that should help to handle $5$ lines in $\mathbb{P}_5$ or to understand $(G(2,V_6))^5/PGL(V_6)$. \begin{corollary}\label{unirationalisation} Five lines in general position in $\mathbb{P}_5$ can be put in the following form: $$\eps_0 \wedge \eps_3,\ \eps_1 \wedge \eps_4,\ \eps_2 \wedge \eps_5,\ (\eps_0+\eps_1+\eps_2)\wedge(\eps_3+\eps_4+\eps_5)$$ $$(-\eps_0+v\eps_4-\eps_5) \wedge(u.\eps_1-\eps_2+a_{1,2,4}\eps_3+ e_1.\eps_4+e_2.\eps_5)$$ for some basis $(\eps_i)_{0\leq i \leq 5}$ of $V_6$, and some complex parameters $u,v,a_{1,2,4},e_1,e_2$. \end{corollary} {\noindent\it Proof: } Let's use again notations of proposition \ref{P4-5sec}. From five general lines in $\mathbb{P}_5$, we obtain a five dimensional subspace $W_5 ^{\vee}$ of $\bigwedge\limits^2 V_6$ containing the corresponding decomposable elements. So choose a general four dimensional vector subspace $W_4 ^{\vee}$ of $W_5 ^{\vee}$, then $(W_4 ^{\vee},W_5 ^{\vee})$ is a general element of the incidence variety $I_{4,5}$. So from Theorem \ref{explicit} and Corollary \ref{bir} the corresponding element of $W_5\otimes \bigwedge\limits^2 V_6$ can be written with notation of definition \ref{defexplicit}: $M_{0123}+w_4.M_4$. So we obtain the proposition. \nolinebreak $\Box$ \subsection{Questions on the magic square} \begin{remark} Let $X$ be a non degenerate subvariety of $\mathbb{P}_{n-1}$. Then the projection of $X$ from a general linear space of dimension $d-2$ is expected to have a finite number $n_{d,X}$ of points of multiplicity $d$ when: $$ d^2+d(\dim(X)-n-1)+n=0.$$ \end{remark} Varieties related to the magic square are famous solutions of this problem for $d=2$ or $d=3$ with $n_{d,X}=1$. For these varieties, what is the number $n_{\frac{n}{d},X}$? For the Veronese surface we have $n_{2,X}\neq n_{3,X}$, but for $\mathbb{P}_2 \times \mathbb{P}_2$, $v_3(\mathbb{P}_1)$, $\mathbb{P}_1 \times \mathbb{P}_1 \times \mathbb{P}_1$ (Cf \cite{H}) we have $n_{d,X}=n_{\frac{n}{d},X}=1$. And now, from proposition \ref{P4-5sec} this equality is also true for $G(2,6)$. \section{Applications }\label{appli} Let $V_{10}$ be a $10$-dimensional vector space over the complex numbers. In this section, we will first explain the relationship between two known constructions associated to the choice of a general element of $\bigwedge\limits^3 V_{10}$. Then we will discuss how the results of the previous section should be related to the symplectic form of the varieties constructed in \cite{D-V}. \subsection{Peskine's example in $\mathbb{P}_9$} This example was constructed by C. Peskine to obtain a smooth non quadratically normal variety of codimension $3$. Let $\mathbb{P}_9$ be a $9$ dimensional projective space over the complex numbers, and denote by $V_{10}$ the vector space $V_{10}=H^0 ({\cal O}_{\mathbb{P}_9}(1))$. Let $\alpha$ be a general element of $\bigwedge\limits^3 V_{10}$, and denote by $\Omega^i_{\mathbb{P}_9}$ the $i$-th exterior power of the cotangent sheaf of $\mathbb{P}_9$. From the identification $\bigwedge\limits^3 V_{10} = H^0 \Omega^2_{\mathbb{P}_9}(3)$, we obtain a skew-symmetric map $M_\alpha$ from $(\Omega^1_{\mathbb{P}_9}) ^{\vee}(-1)$ to $\Omega^1_{\mathbb{P}_9}(2)$ and an exact sequence: $$ \begin{CD} 0@>>> {\cal O}_{\mathbb{P}_9}(-3) @>>> (\Omega^{1}_{\mathbb{P}_9}) ^{\vee}(-1)@>M_{\alpha}>>\Omega^1_{\mathbb{P}_9}(2)@>>> I_{Y_\alpha}(4) @>>> 0 \end{CD} $$ where $I_{Y_\alpha}$ is the ideal of the smooth variety of dimension $6$ defined by the $8$ by $8$ pfaffians of $M_{\alpha}$. The following proposition is directly deduced from the previous exact sequence. \begin{proposition} The variety $Y_{\alpha}$ is such that $h^1(I_{Y_{\alpha}}(2))=1$ and its canonical sheaf is $\omega_{Y_{\alpha}}={\cal O}_{Y_{\alpha}}(-3)$. \end{proposition} \subsection{Debarre-Voisin's manifold as a parameter space} Denote by $G(6,V_{10} ^{\vee})$ the Grassmannian of $6$ dimensional subspaces of $V_{10} ^{\vee}$. Let $K_6$ (resp. $Q_4$) be the tautological subbundle (resp. quotient bundle). For any $p\in G(6,V_{10} ^{\vee})$, the corresponding $5$-dimensional projective subspace of $\mathbb{P}_9$ will be denoted by $\kappa_p$. \medskip Debarre and Voisin proved in \cite{D-V} the following: \begin{theorem}(\cite{D-V} Th 1.1). Let $\alpha$ be a general element of $\bigwedge\limits^3 V_{10} = H^0(\bigwedge\limits^3 K_6 ^{\vee})$. The subvariety $Z_{\alpha}$ of $G(6,V_{10} ^{\vee})$ defined by the vanishing locus of the section $\alpha$ of $\bigwedge\limits^3 K_6 ^{\vee}$ is an irreducible hyper-K\"ahler manifold of dimension $4$ and second betti number $23$. \end{theorem} We can now remark the following relation between $Y_{\alpha}$, $Z_{\alpha}$ and Palatini threefolds: \begin{proposition} Let $p$ be a general element of $Z_{\alpha}$. The scheme defined by the intersection $Y_{\alpha}\cap \kappa_p$ is a Palatini threefold. \end{proposition} {\noindent\it Proof: } The restriction of $\Omega^1_{\mathbb{P}_9}(1)$ to $\kappa_p$ is $\Omega^1_{\kappa_p}(1)\oplus 4{\cal O}_{\kappa_p}$. The vanishing of the restriction of $\alpha$ to $\kappa_p$ implies that the restriction of $M_\alpha$ to $\kappa_p$ is: $\left(\begin{array}[c]{cc} 0 & \alpha_p\\ -\t{\alpha_p} & \beta \end{array}\right)$ with respect to the direct sums: $ (\Omega^1_{\kappa_p}) ^{\vee}(-1)\oplus 4 {\cal O}_{\kappa_p} \to (\Omega^1_{\kappa_p})(2) \oplus 4{\cal O}_{\kappa_p}(1)$. So the ideal generated by the pfaffians of size $8$ of this map is also the ideal generated by the maximal minors of $\alpha_p: 4 {\cal O}_{\kappa_p} \to (\Omega^1_{\kappa_p})(2)$. In conclusion the scheme defined by the intersection $Y_{\alpha}\cap \kappa_p$ is a Palatini threefold as in remark \ref{resolpala}. \nolinebreak $\Box$ \bigskip Moreover, the following construction globalize definition \ref{palatini} and the pfaffian cubic surface over $Z_{\alpha}$. \begin{remark}\label{Zreseau} The restriction of the bundle $\bigwedge\limits^2 K_6 ^{\vee} \otimes Q_4 ^{\vee}$ to $Z_{\alpha}$ has a non trivial section. It gives an injective map: $$ (Q_4)_{|Z_{\alpha}} \longrightarrow (\bigwedge\limits^2 K_6 ^{\vee})_{|Z_{\alpha}} $$ \end{remark} {\noindent\it Proof: } The section $\alpha$ of $(\bigwedge\limits^3 K_6 ^{\vee})$ gives a map from $K_6$ to $(\bigwedge\limits^2 K_6 ^{\vee})$. But the restriction of this map to $Z_{\alpha}$ is zero, so it induces a map from the quotient $(Q_4)_{|Z_{\alpha}}$ to $(\bigwedge\limits^2 K_6 ^{\vee})_{|Z_{\alpha}} $. The injectivity of this maps of ${\cal O}_{Z_{\alpha}}$-modules follows from the assumption that $\alpha$ is general. \nolinebreak $\Box$ \subsection{Conjectures on the symplectic form on $Z_{\alpha}$} \begin{remark}\label{5P1} Let $p$ be a general element of $Z_{\alpha}$. The tangent space $\mathcal{T}_{(Z_{\alpha},p)}$ to $Z_{\alpha}$ at $p$ contains a canonical set of $5$ vector spaces of dimension $2$. \end{remark} {\noindent\it Proof: } Let $p$ be a general point of $Z_{\alpha}$. From remark \ref{Zreseau}, the fiber $Q_{4,p}$ is a $4$-dimensional subspace of $(\bigwedge\limits^2 K_{6,p} ^{\vee})$. From proposition \ref{P4-5sec}, we obtain in $K_{6,p} ^{\vee}$, a canonical set of five vector subspaces $(L_i)_{0\leq i \leq 4}$ of dimension $2$ such that $\bigoplus\limits_{0\leq i \leq 4} \bigwedge\limits^2 L_i$ contains $Q_{4,p}$. So the restriction of the map: \begin{equation} \label{tens21vers3} m_{21}: \bigwedge\limits^2 K_{6,p} ^{\vee} \otimes K_{6,p} ^{\vee} \to \bigwedge\limits^3 K_{6,p} ^{\vee} \end{equation} gives the following commutative diagram of exact sequences: $$ \begin{CD} 0 @>>> \mathcal{T}_{(Z_{\alpha},p)} @>>> Q_{4,p} \otimes K_{6,p} ^{\vee} @>>> \bigwedge\limits^3 K_{6,p} ^{\vee} @>>> 0 \\ @. @. @VVV @VVV \\ @. @. (\bigoplus\limits_{0\leq i \leq 4} \bigwedge\limits^2 L_i )\otimes K_{6,p} ^{\vee} @>>> \bigwedge\limits^3 K_{6,p} ^{\vee} @>>>0 \end{CD} $$ where the vertical maps are injectives and the first row is the normal sequence of $Z_{\alpha}$ in $G(6,V_{10} ^{\vee})$ at the point $p$. Now remark that $m_{21}$ vanishes on each $\bigwedge\limits^2 L_i \otimes L_i$ because $L_i$ has dimension $2$. So we can identify the kernel of the second row of the previous diagram with the $10$-dimensional vector space $ \bigoplus\limits_{0\leq i \leq 4}\bigwedge\limits^2 L_i \otimes L_i$, and we obtain an injection: $$\mathcal{T}_{(Z_{\alpha},p)} \hookrightarrow \bigoplus\limits_{0\leq i \leq 4}\bigwedge\limits^2 L_i \otimes L_i .$$ So in general, the kernel of each projection $ \mathcal{T}_{(Z_{\alpha},p)} \to \bigwedge\limits^2 L_i \otimes L_i$ gives a $2$ dimensional vector subspace of $ \mathcal{T}_{(Z_{\alpha},p)}$. \nolinebreak $\Box$ \bigskip Now we can remark that five points of $G(2,\mathcal{T}_{Z_{\alpha},p})$ should define an hyperplane $\gamma$ in $\bigwedge\limits^2 \mathcal{T}_{Z_{\alpha},p}$. Some random examples with \cite{Macaulay2} let us expect that the ideal of these five lines $(l_i)$ in $\mathbb{P}(\mathcal{T}_{Z_{\alpha},p})$ is given by the maximal minors of the map: $$ K_{6,p} \otimes {\cal O}_{\mathbb{P}(\mathcal{T}_{Z_{\alpha},p})} \to Q_{4,p}\otimes {\cal O}_{\mathbb{P}(\mathcal{T}_{Z_{\alpha},p})}(1) $$ obtained from the inclusion of the tangent space to $Z_{\alpha}$ in the tangent space to $G(6,V_{10} ^{\vee})$. But if the alternate form $\gamma$ was degenerated, its kernel would give a line in $\mathbb{P}(\mathcal{T}_{Z_{\alpha},p})$ intersecting each $l_i$. But a variety defined by quartic hypersurfaces can't have a $5$-secant line, so we can expect the following: \begin{conjecture} The five vector spaces of dimension $2$ canonically defined in the remark \ref{5P1} are maximal isotropic subspaces for the symplectic form on $\mathcal{T}_{Z_{\alpha}}$ constructed by Debarre and Voisin. \end{conjecture}
\section{Introduction} \label{} The use of isotopes as geochemical tracers depends upon the existence of reliable $P-T$-dependent equilibrium isotope fractionation data between solids, fluids and melts. The common method used for determination of such data is to conduct experiments, where the phases of interest are equilibrated at a range of different conditions and individually measured for their equilibrium isotopic compositions. Recently, with the development of computational methods, software and increase in hardware performance it is also possible to simulate and compute the isotope fractionation with {\it ab initio} methods. These computational considerations complement the experimental effort and provide information on the mechanisms governing the equilibrium isotope fractionation processes on the atomic scale. By establishing an efficient computational approach for materials at high $P$ and $T$ and testing its reliability by computing Li isotope fractionation between minerals and aqueous fluids, \citet{KJ11} have shown that {\it ab initio} methods can provide a reliable estimation of equilibrium isotope fractionation factors at an accuracy level comparable to experiments. Motivated by the encouraging results for the fractionation of lithium isotopes we applied our new method to study boron isotope fractionation. The main goals of the present study are to make theoretical predictions and obtain a better understanding of the B isotope fractionation process between tourmaline, B-bearing mica and fluids at various pressures and temperatures, to compare these data to results from recent {\it in situ} experimental studies \citep{WM05,MW08} and measurements of isotopic signatures in natural samples \citep{KM11,M05,H02} and to investigate the underlying mechanisms driving the boron isotope fractionation processes between the considered materials. With two stable isotopes $^{10}$B and $^{11}$B, of relatively large mass difference of about $10\,\%$, boron isotopes strongly fractionate during geological processes, thus leading to natural $\rm\delta^{11}B$-variations ranging from $-30$ to $\rm+60\,\permil$ \citep{B93}. Therefore, B isotopes are ideal for the distinction of different geological environments and for quantifying mass transfer processes, e.g. in the range of subduction zones. First-order criteria driving isotope fractionation in Earth materials are differences in coordination and in the bonding environments of coexisting phases. The lighter isotope usually preferentially occupies the higher coordinated site, which is generally accompanied with a longer cation-anion bond length and weaker bond strength \citep{SMH09,WJ11}. Tourmaline has an extensive chemical variability and is the most widespread borosilicate in various rocks over a large range of bulk compositions. It has a large $P-T$-stability which ranges from surface conditions to high pressures and temperatures of at least 7.0 GPa and about $\rm 1000^{o}C$ as determined experimentally for dravitic tourmaline \citep{K95}. Most tourmaline minerals contain 3 boron atoms per formula unit (pfu) with B in trigonal planar coordination (B$^{[3]}$). High pressure combined with an Al-rich environment can lead to the formation of olenitic tourmaline with significant amounts of excess B substituting for Si at the tetrahedral site (B$\rm^{[4]}$). The highest amounts of up to 1.2 B$^{[4]}$ pfu have been found in olenitic tourmaline from Koralpe, Austria \citep{EP97}. Al-rich tourmaline with up to 2.2 B$^{[4]}$ pfu was synthesized experimentally at $\rm 2.5\,GPa$, 600$^{\rm o}$C \citep{SW00}. The maximum possible amount of B$\rm ^{[4]}$ in olenitic tourmaline is limited to three B$\rm^{[4]}$ pfu, due to structural and crystal chemical reasons. As tourmaline is not stable at basic pH \citep{ML89}, the dominant B-species of fluids coexisting with tourmaline is $\rm B(OH)_3$. Therefore, due to the absence of change in boron coordination, B-isotopic fractionation between B$\rm^{[4]}$-free tourmaline and fluid ($\rm \Delta^{11}B$(tour-fl)) should be small. However, at $0.2\rm\,GPa$, the experimentally determined $\rm \Delta^{11}B$(tour-fl)-values of $-2.5\pm0.4\,\permil$ at 400$\rm^{o}C$ and $-0.4 \pm 0.4\,\permil$ at 700$\rm^{o}C$ \citep{MW08} suggest small but significant differences in the B-O(H) bond strength between tourmaline and the neutral fluids. Incorporation of $\rm B^{[4]}$ into tourmaline is expected to significantly increase the fractionation of boron isotopes between the mineral and aqueous fluid (i.e. increase of $|\rm \Delta^{11}B|$(tour-fl)). In this contribution we present calculated $P-T$-dependent data on B-isotope fractionation between B$\rm ^{[4]}$-bearing tourmaline and fluid, which so far are not available from experiments. Despite of its low boron content of up to maximum values of $\rm 270\,ppm$ \citep{DH93}, potassic white mica is probably the main host of boron in metasedimentary and metabasaltic blueschists and eclogites, because of its high modal abundance and B-incompatibility in all other stable minerals of these rocks \citep{BRS98}. During subduction the modal amount of mica continuously decreases by dehydration reactions and the chemistry of residual micas is shifted towards phengitic compositions. Phengitic mica has an extended stability, extending as deep as $300\rm\,km$ within cold subduction zones \citep{S96}. In contrast to most tourmalines, boron is tetrahedrally coordinated in mica, where it substitutes for aluminum. Due to the coordination change from mostly three-fold coordinated in near-neutral fluids \citep{STH05} to B$\rm^{[4]}$ in mica, B-isotopic fractionation between B-bearing mica and such fluids is much larger than for tourmaline -- fluid. $\rm \Delta^{11}B$(mica-fl)-values determined experimentally at $P=3.0\rm\, GPa$, are $\rm-10.9\pm1.3\,\permil$ at 500$\rm^{o}C$ and $\rm-7.1\pm0.5\,\permil$ at 700$\rm ^{o}C$ \citep{WM05}. Such a strong B-isotope fractionation and its pronounced $T$-dependence, in combination with the continuous dehydration of micas during ongoing subduction and boron transport via fluids into mantle wedge regions of arc magma-formation, probably determines the boron variations and B-isotopic signatures in volcanic arcs \citep{WM05}. Using {\it in situ} Raman spectroscopic measurements of near-neutral B-bearing fluids, \citet{STH05} observed a significant amount of 4-fold B-species at high $P$ and $T$. The abundance of these species increases with temperature and pressure and at $T\rm=800\,K$ and $P\rm=1.9\,GPa$ there should be a considerable amount ($\rm 15-30\%$) of $\rm B^{[4]}$ species in the fluid. Such a significant amount of tetrahedrally coordinated B-species in high temperature and pressure fluids should affect solid--fluid B-isotopic fractionation, which we investigate in our calculations. In the light of this, we also discuss recently determined B-isotope data from coexisting natural tourmaline and B-bearing mica \citep{KM11,M05,H02}, which show slight inconsistency with {\it in situ} measurements of \citet{WM05} and \citet{MW08}. Reliable {\it ab initio} computational methods to predict isotope fractionation factors have been established recently. Several groups have proved that such calculations can contribute towards understanding geochemical mechanisms responsible for production of isotope signatures \citep{D97,YMMW01,S04,DK08,HS08,ML09,ML07,SMH09,Z09,HS10,RC10,RB10,Z09,Z10,KJ11}. The majority of these works, however, concentrate on the computation of the stable isotope fractionation between various, mostly simple crystalline minerals, and the aqueous solutions are usually computed using an isolated cluster containing fractionating species and a hydration shell (e.g. \citet{Z10,Z09,HS10,RC10,DK08,HS08,RB10,Z05}). However, aqueous solutions at high pressure and temperature must be computed with caution as discussed in \citet{KJ11}. This is because the distribution of cation coordination and cation-oxygen bond lengths that affect the isotope fractionation \citep{BM47} is driven by the dynamics of the system and change under compression \citep{JW09,WJ11,KJ11}. The only recent {\it ab initio} work, besides \citet{KJ11}, that accounts for the dynamical effects on the isotope fractionation in fluid is by \citet{RB07} who considered boron equilibrium isotope fractionation between $\rm B(OH)_3$ and $\rm B(OH)_4^-$ species in aqueous solution. They performed {\it ab initio} molecular dynamics simulations of these fluids and attempted to use the vibrational density of states, derived through the Fourier transform of the velocity auto-correlation function, as an input for the calculation of the $^{11}$B/$^{10}$B isotope fractionation coefficient. However, the resulting fractionation factor $\alpha=0.86$ happened to be much lower than the experimental value of $\alpha=1.028$. Interestingly, the discrepancy between experiment and theory is cured by quenching the selected configurations along the molecular dynamics trajectory and computing the harmonic frequencies. The fractionation factor derived using these frequencies exactly reproduces the experimental value. In our approach both solids and fluids are treated as extended systems by application of periodic boundary conditions in all three spatial directions, which is crucial to model high pressure materials. Large enough supercells are chosen to avoid significant interaction between atoms and their periodic images, as well as to reduce the number of k-points and q-vectors for sampling the Brillouin zones and the phonon spectra of the crystals, respectively (for a liquid or fluid, both the Brillouin zone and phonons are not defined). In our investigation we use cells of at least $7\rm\,\AA$ width in each spatial dimension. A representative statistical sampling of the fluid structure is obtained by performing Car-Parrinello molecular dynamics simulations \citep{CPMD1}. For the calculation of the isotope fractionation factors in fluids, several random snapshots from the simulation runs are chosen. The force constants acting on the fractionating element and the resulting fractionation factors are then obtained for each ionic configuration, and the relevant fractionation factor for the boron species in the fluid is computed as an average over the whole set of considered geometries. In \citet{KJ11} it was shown that in line with the \citet{BM47} approximation, considering the force constants acting on the fractionating atom only leads to a satisfactory estimation of the Li isotope fractionation factors for high temperature fluids and minerals. Here, we will show that approximating the vibrational spectrum by the three pseudofrequencies derived from the force constants allows for further improvement of the accuracy of the predicted isotope fractionation factors, especially at lower temperatures. In this contribution we present the theoretical prediction of B isotope fractionation factors between B bearing aqueous fluids and solids, specifically tourmaline and B-muscovite, for which experimental data and measurements on natural samples are available for comparison \citep{WM05,MW08,KM11}. We will show that the application of {\it ab initio} methods to B-bearing crystalline solids and fluids not only provides unique insight into the mechanisms driving equilibrium B-isotope fractionation on the atomic scale, but helps in proper interpretation of the data. \begin{table*}[t] \caption{The computed and measured vibrational frequencies for $\rm H_3BO_3$ and $\rm H_4BO_4^-$ molecules reported for two different boron isotopes B$^{11}$/B$^{10}$. Only the frequencies affected by the isotopic substitution are reported. The units are $\rm cm^{-1}$.} \label{T1a} \centering \begin{tabular}{lcccccc} \hline\hline method & $\rm H_3BO_3$ & & $\rm H_4BO_4^-$ & & & \\ \hline theoretical: & & & & & & \\ BLYP (this work) & 656/680 & 1366/1408 & 793/807 & 870/886 & 1051/1063 & 1132/1154 \\ BLYP $^1$ & 642/667 & 1390/1437 & 807/821 & 916/947 & 1020/1025 & 1156/1179 \\ HF/6-31G$^*$ $^2$ & 736/764 & 1562/1615 & & & & \\ HF/aug-cc-pVDZ$^*$ $^3$ & \hspace{4mm}-/754 & \hspace{6mm}-/1531 & & & & \\ BP86/aug-cc-pVDZ$^*$ $^3$ & \hspace{4mm}-/655 & \hspace{6mm}-/1412 & & & & \\ B3LYP/aug-cc-pVDZ$^*$ $^3$ & \hspace{4mm}-/684 & \hspace{6mm}-/1447 & & & & \\ MP2/aug-cc-pVDZ$^*$ $^3$ & \hspace{4mm}-/684 & \hspace{6mm}-/1435 & & & & \\ CCSD(T)/aug-cc-pVDZ$^*$ $^3$ & \hspace{4mm}-/684 & \hspace{6mm}-/1438 & & & & \\ experimental: & & & & & & \\ solution $^4$ & 632/666 & 1412/1454 & & 937/975 & & \\ vapour $^5$ & 674/700 & 1429/1477 & & & & \\ solution $^6$ & 639/668 & 1428/1490 & & 947/-\hspace{4mm} & & \\ vapour $^7$ & 666/692 & 1415/1472 & & & & \\ vapour $^8$ & 675/701 & 1426/1478 & & & & \\ solution and vapour $^9$ & 639-675/- & 1421-1450/- & & 935-958/- & & \\ \hline \end{tabular} \\ References: $^1$\citet{Z05}, $^2$\cite{LT05}, $^3$\cite{RB10}, $^4$\cite{SV05}, $^5$\cite{G91}, $^6$\cite{O00,LT05}, $^7$\cite{AB92}, $^8$\cite{OY88}, $^9$\cite{Z05} and references hereafter. \end{table*} \begin{table*}[t] \caption{$1000(\beta-1)$ factors for isolated $\rm H_3BO_3$ and $\rm H_4BO_4^-$ molecules at $T\rm=300\,K$ obtained using three methods: (1) the full frequency spectrum and Equation \ref{beta}, (2) {\it the single atom approximation} of \citet{KJ11}, (3) {\it the single atom approximation} with {\it the pseudofrequencies} and equation \ref{beta}. The units are $\permil$. The $\Delta\beta/\beta_{\rm BM}$ is the check of condition given by Eq. \ref{TESTP} and indicates the improvement of the method (3) over method (2) expressed in $\%$.} \label{T1} \centering \begin{tabular}{lrcccrccc} \hline\hline T (K) & meth. (1) & meth. (2) & meth. (3) & $\Delta\beta/\beta_{BM}\,[\%]$ & meth. (1) & meth. (2) & meth. (3) & $\Delta\beta/\beta_{BM}\,[\%]$\\ \hline & $\rm H_3BO_3$ & & & & $\rm H_4BO_4^-$ & & & \\ 300 & 211.3 & 283.1 & 225.1 & 20.5 & 167.3 & 203.4 & 175.2 & 13.9 \\ 600 & 64.1 & 70.8 & 65.6 & 7.3 & 48.1 & 50.9 & 48.6 & 4.5 \\ 800 & 37.6 & 39.8 & 38.1 & 4.3 & 27.8 & 28.6 & 27.9 & 2.5 \\ 1000 & 24.6 & 25.4 & 24.7 & 2.8 & 18.0 & 18.3 & 18.0 & 1.7 \\ & Dravite & & & & Boromuscovite 1M & & & \\ 300 & 202.8 & 267.7 & 216.0 & 19.3 & 139.1 & 160.8 & 142.2 & 11.6 \\ 600 & 60.7 & 66.9 & 62.4 & 6.7 & 38.6 & 40.2 & 38.8 & 3.5 \\ 800 & 35.5 & 37.6 & 36.1 & 4.0 & 22.1 & 22.6 & 22.2 & 1.8 \\ 1000 & 23.2 & 24.1 & 23.5 & 2.5 & 14.3 & 14.5 & 14.3 & 1.4 \\ \hline \end{tabular} \\ \end{table*} \section{Computational approach} \subsection{Theoretical model \label{TM}} \subsubsection{The single atom approximation: \citet{BM47} approach} Mass-dependent equilibrium isotope fractionation is driven by the change in molecular and crystalline vibration frequencies resulting from the different masses of the isotopes. The fractionation between species and an ideal monoatomic gas is called the $\beta$ factor. In the harmonic approximation it is given by the formula \citep{BM47,U47,CCH01}: \begin{equation} \beta=\prod_{i=1}^{N_{dof}}\frac{u_{i}^{*}}{u_{i}}\exp{\left[ \frac{(u_{i}-u_{i}^*)}{2}\right]}\frac{1-\exp(-u_i)}{1-\exp(-u_i^*)}, \label{beta}\end{equation} where $u={h\nu_i}/{k_BT}$, $h$ is the Planck constant, $\nu_i$ is the vibrational frequency of the $i$-th degree of freedom, $k_B$ is the Boltzmann constant, $N_{dof}$ is the number of degrees of freedom, which for $N$ being the number of atoms in the considered system (molecule, mineral or fluid) is equal to $3N-5$ for a diatomic molecule, $3N-6$ for multiatomic molecules and $3N$ for crystals, and a star symbol marks the heavier isotope. The fractionation factor between two substances A and B, $\alpha_{A-B}$ is computed as the ratio of the relevant $\beta$ factors, which is well approximated by the differences in the $\beta$ factors: \begin{equation} \alpha_{A-B}=\beta_A/\beta_B,\,\Delta_{A-B}\cong 1000\ln\beta_A-1000\ln\beta_B [\permil].\end{equation} The calculation of the $\beta$ factor requires only knowledge of the vibrational properties of the considered system computed for the two different isotopes. However, computation of the whole vibrational spectra of complex, multiparticle minerals or fluids requires substantial computational resources and is currently limited to systems containing a few dozens of atoms or less. In our recent work \citep{KJ11} we proposed to use an efficient method for computing the high temperature isotope fractionation factors between complex materials such as fluids and crystalline solids, which requires the knowledge of the force constants acting upon the fractionating element only. The $\beta$ factor (Eq. \ref{beta}) can be then approximated by \citep{BM47,KJ11}: \begin{equation} \beta\simeq 1+\sum_{i=1}^{N_{dof}}\frac{u_i^2-u_i^{*2}}{24}=1+\frac{\Delta m}{m m^*}\frac{\hbar^2}{24 k_B^2T^2}\sum_{i=1}^3 A_i, \label{BAPX} \end{equation} where $A_i$ are the force constants acting on the isotopic atom in the three perpendicular spatial directions (x, y and z), $\Delta m=m^*-m$, where $m$ and $m^*$ are the masses of the lighter and heavier isotopes of the fractionating element. As the computation of the $\beta$ factors from formula \ref{BAPX} requires the knowledge of properties of the fractionating element only we will call such an approach {\it the single atom approximation} throughout the paper. The validity criteria restricts the usage of the formula to frequencies $ \nu\,{\rm[cm^{-1}]}\lesssim1.39\,T\rm{[K]}$ (assuming $u<2$, see Fig. 1 of \citet{BM47}). We are interested in temperature range $800$-$1000\rm\,K$. The highest vibrational frequency of the modes involving movement of B atoms for $\rm H_3BO_3$ is $\sim1400\,\rm cm^{-1}$ and of $\rm H_4BO_4^{-1}$ is $<1200\,\rm cm^{-1}$ (Table \ref{T1a}). In case of $\rm H_3BO_3$, the single atom approximation may produce an error of $2.2\, \permil$ in the $\beta$ factor at $T\rm=800\,K$. The relevant error for $\rm H_4BO_4^{-1}$ is $0.8\, \permil$ (Table \ref{T1}). A further improvement to the method is therefore desired. \subsubsection{The single atom approximation with pseudofrequencies: our improvement \label{SAA}} We will show that the error of {\it the single atom approximation} can be substantially reduced if one uses the three frequencies $\bar{\nu_i}$ derived from the force constants acting on the fractionating element ($\bar{\nu}_i^2=A_i/4\pi^2m$). We call them {\it``pseudofrequencies"}, and compute the $\beta$ factors using formula \ref{beta}. In the following we present the formal justification of such an approach. According to \citet{BM47}, equation \ref{beta} for small $\Delta u_i=u_i-u_i^*$ reduces to (\citet{BM47}, Eq. 11a): \begin{equation} \beta=1+\sum_{i=1}^{N_{dof}}\left (\frac{1}{2}-\frac{1}{u_i}+\frac{1}{\exp(u_i)-1}\right )\Delta u_i. \label{BEXP}\end{equation} The Taylor expansion of the function appearing under the summation sign is: \begin{equation*} G(u)=\frac{1}{2}-\frac{1}{u}+\frac{1}{\exp(u_i)-1} \end{equation*} \begin{equation}=\frac{u}{12}-\frac{u^3}{720}+\frac{u^5}{30240}-\frac{u^7}{1209600}+... .\end{equation} When we consider just the first term of the expansion the $\beta$ factor is: \begin{equation*} \beta=1+\sum_i^{N_{dof}} \frac{u_i}{12}\Delta u_i \end{equation*} \begin{equation}=1+\sum_{i=1}^{N_{dof}} \frac{\Delta u_i^2}{24}=1+\sum_{i=1}^{N_{dof}} \frac{u_i^2-u_i^{*2}}{24},\end{equation} which is exactly equation \ref{BAPX}. \begin{table}[t] \caption{The three pseudofrequencies of the B atom derived for two selected molecules and two crystalline solids. The units are $\rm cm^{-1}$.} \label{T1b} \centering \begin{tabular}{lccc} \hline\hline species & & & \\ \hline $\rm H_3BO_3$ & 644/675 & 1129/1184 & 1130/1185 \\ dravite & 618/648 & 1085/1138 & 1120/1174 \\ $\rm H_4BO_4^-$ & 809/848 & 845/886 & 872/915 \\ boromuscovite 1M & 722/757 & 723/758 & 801/840 \\ \hline \end{tabular} \end{table} \begin{figure*}[t] \includegraphics[angle=270,width=3.5in]{fig1.eps} \caption{ $\beta$ factors for isolated $\rm H_3BO_3$ and $\rm H_4BO_4^-$ molecules as well as dravite and boromuscovite 1M crystalline solids. The lines represent the results obtained using: (1) the full frequency spectrum and Eq. \ref{beta} (solid), (2) {\it the single atom approximation } of \citet{BM47,KJ11}, Eq. \ref{BAPX} (dotted) and (3) the method described in the text with {\it the pseudofrequencies} and Eq. \ref{beta} (dashed lines). \label{F1}} \end{figure*} \begin{table*}[t] \caption{The lattice parameters of the investigated B-bearing crystalline solids. $N_{atoms}$ is the number of atoms in the modeled supercell.} \label{T2} \centering \begin{tabular}{lcccccc} \hline\hline & boromuscovite 1M$^1$ & boromuscovite 2M1$^1$ & dravite$^2$ & olenite$^3$ \\ \hline a (\AA) & 10.204 & 10.180 & 15.945 & 15.5996 \\ b (\AA) & 8.788 & 8.822 & 15.945 & 15.5996 \\ c (\AA) & 10.076 & 19.8189 & 7.210 & 7.0224 \\ $\alpha$ ($^\circ$) & 90 & 90 & 90 & 90 \\ $\beta$ ($^\circ$) & 101.23 & 95.62 & 90 & 90 \\ $\gamma$ ($^\circ$) & 90 & 90 & 90 & 120 \\ $N_{atoms}$ & 84 & 168& 163 & 162\\ \hline \end{tabular} \\ References: $^1$\cite{LH95}, $^2$\cite{EM10}, $^3$\cite{MB02} \end{table*} Let us consider the Taylor expansion of the different estimations of $\beta$ factors. Equation \ref{BEXP} then reads: \begin{equation} \beta=\beta_{exact}=1+\sum_{i=1}^{N_{dof}}\left (\frac{u_i}{12}-\frac{u_i^3}{720}+... \right )\Delta u_i \label{B11}.\end{equation} The \citet{BM47} approximation given by equation \ref{BAPX} reads: \begin{equation} \beta\sim\beta_{BM}=1+\sum_{i=1}^{N_{dof}}\frac{u_i}{12}\Delta u_i \label{B111},\end{equation} and the proposed approximation based on pseudofrequencies is: \begin{equation} \beta\sim\beta_{pseudo}=1+\sum_{i=1}^3\left (\frac{\bar{u}_i}{12}-\frac{\bar{u}_i^3}{720}+... \right )\Delta \bar{u}_i. \label{B22}\end{equation} In the above equation $\bar{u_i}=h\bar{\nu_i}/k_BT$. Next we check how this approximation compares to the \citet{BM47} approximation given by equations \ref{BAPX} and \ref{B111}. In order to make a comparison we derive the differences between the two approximate expressions $\beta_{BM}$, $\beta_{pseudo}$ and the exact one $\beta_{exact}$ (Eq. \ref{B11}). In the case of the \citet{BM47} approximation we have: \begin{equation*} \Delta\beta_{BM}=\beta_{BM}-\beta_{exact}=\left (\sum_i^{N_{dof}}\frac{u_i}{12} -\sum_i^{N_{dof}} \left (\frac{u_i}{12}-\frac{u_i^3}{720}+... \right)\right)\Delta u_i \end{equation*} \begin{equation} =\sum_i^{N_{dof}}\left ( \frac{u_i^3}{720}-... \right )\Delta u_i\end{equation} and having from equations \ref{BAPX} and \ref{B111} that \begin{equation} \sum_i^{N_{dof}}\frac{u_i}{12}\Delta u_i=\sum_i^{3}\frac{\bar{u_i}}{12}\Delta \bar{u_i} \label {BMBM} \end{equation} in the case of the proposed approximation we get: \begin{equation*} \Delta\beta_{pseudo}=\beta_{pseudo}-\beta_{exact}=\sum_{i=1}^3\left (-\frac{\bar{u}_i^3}{720}+...\right )\Delta \bar{u}_i \end{equation*} \begin{equation*} +\sum_i^{N_{dof}}\left (\frac{u_i^3}{720}-...\right)\Delta u_i=\end{equation*} \begin{equation} =\sum_{i=1}^3\left (-\frac{\bar{u}_i^3}{720}+...\right )\Delta \bar{u}_i+\Delta\beta_{BM} \label{BPS}\end{equation} Because relation $G(u)<\frac{u}{12}$ holds for any $u$ (see \citet{BM47}, Fig. 1), the function \begin{equation} \sum_{i=1}^3\left (-\frac{\bar{u}_i^3}{720}+...\right )\Delta u_i=\sum_{i=1}^3\left ( G(\bar{u_i})-\frac{\bar{u_i}}{12}\right ) \Delta \bar{u_i}<0\end{equation} and $\Delta\beta_{pseudo}<\Delta\beta_{BM}$. On the other hand the expression for $\Delta\beta_{pseudo}$ is given by a difference of the higher order terms of the Taylor expansions of the two expressions for the $\beta$ factor. In the considered cases the values of pseudofrequencies are similar to the real frequencies that are affected upon B isotope substitution in a given B-bearing system. This can be seen by comparing the pseudofrequencies computed for the selected cases of B-bearing molecules and crystalline solids considered here and reported in Table \ref{T1b} with the real frequencies given in Table \ref{T1a}. This indicates that the two terms of opposite signs in Eq. \ref{BPS} should be similar in value and cancel out to a great extent, so $|\Delta\beta_{pseudo}|<<\Delta\beta_{BM}$. Therefore the approach proposed here to compute the $\beta$ factor based on pseudofrequencies and Eq. \ref{beta} should give a better approximation to the exact $\beta$ factors than equation \ref{BAPX}, which we will show in section \ref{TEST}. The difference to the \citet{BM47} approximation is given by: \begin{equation} \Delta\beta=1+\sum_{i=1}^3\frac{\bar{u_i}}{12} \Delta \bar{u_i}-\beta_{pseudo}\end{equation} and can be easily computed for any considered system. We assume that the pseudofrequency-based approach to the computation of $\beta$ factors is applicable if it is just a correction to equation \ref{BAPX}, i.e. when: \begin{equation} \Delta\beta<<\sum_{i=1}^3\frac{\bar{u_i}}{12} \Delta \bar{u_i}. \label{TESTP} \end{equation} \begin{figure*}[t] \includegraphics[angle=270,width=3.5in]{fig2.eps} \caption{Left panel: $\beta$ factors for $\rm H_3BO_3$ (solid line) and $\rm H_4BO_4^-$ (dashed lines) in the gas phase. Right panel: fractionation factors between $\rm H_3BO_3$ and $\rm H_4BO_4^-$ in gas phase. The thick lines represent our results, while the thin lines represent the results obtained using frequencies of \citet{Z05}. \label{F6}} \end{figure*} We also note that the proposed approach satisfies the Redlich-Teller product rule \citep{R35} when the mass of considered isotope $m$ is much smaller than the mass of the whole considered system $M$, namely, \begin{equation} \prod_{i=1}^{N_{dof}}\frac{u_i^*}{u_i}=\left(\frac{M^*}{M}\frac{m}{m^*}\right)^{3/2}\approxeq\left(\frac{m}{m^*}\right)^{3/2}= \prod_{i=1}^{3}\frac{\bar{u}_i^*}{\bar{u}_i}.\end{equation} We notice that we could force the strict conservation of the Redlich-Teller product by just adjusting the ratios of $\bar{u_i}/\bar{u_i^*}$. However, such a modification would not preserve relation \ref{BMBM} and the pseudofrequency approach would not recover exactly the high temperature limit (Eq. \ref{BAPX}) of the exact solution (Eq. \ref{beta}), which is a more important constraint to fulfill strictly by the proposed approximation. \subsection{Representation of solids} In this paper we investigate the boron isotope fractionation between dravite, olenite and boromuscovite minerals and aqueous fluids. The solids were represented by large cells containing at least 84 atoms. The number of atoms used in the crystal calculations together with the lattice parameters of modeled crystals are summarized in Table \ref{T2}. The lattice parameters and chemical compositions of the modeled crystalline solids are the experimental values measured at ambient conditions found in the literature. Dravite is the crystalline solid which was used in the experiments on tourmaline by \citet{MW08}. The chemical composition of the supercell used in the investigation is $\rm Na_{3} Mg_{9} Al_{18} (Si_{18}O_{54})(B^{[3]}O_{3})_{9}(OH)_{12}$ with structural data of \citet{MB02}. Olenite can contain B in both trigonal and tetragonal sites. The modeled structure is that of \citet{EM10}. The chemical composition of the unit cell used in the investigation is $\rm Na Al_{3} Al_{6} (Si_{4}B_{2}^{[4]}O_{18})(B^{[3]}O_{3})_{3}(OH)_{3}O$. For boromuscovite, the 1M and 2M1 crystal structures of \citet{LH95} were used. In the isotope fractionation experiments of \citet{WM05} boromuscovite forms two polytypes, 1M and 2M1, with relative abundances of $10\%$ and $90\%$ respectively. In boromuscovite B occupies the 4-fold coordinated site occupied mainly by Si atoms. The constructed model constitutes a 2x1x1 supercell of elementary chemical composition $\rm K Al_{2} (B^{[4]} Si_{3} O_{10})(OH)_{2}$. \subsection{Representation of aqueous solution} The aqueous solution was represented by a periodically repeated box containing up to 64 water molecules and one $\rm H_3BO_3$ or $\rm H_4BO_4^-$ molecule. The pressure and temperature conditions were chosen to be close to the experimental conditions of \citet{WM05} and \citet{MW08}. The pressure of the aqueous solution for a given temperature and volume was calculated according to the equation of state of \citet{WP02}. The {\it ab initio} molecular dynamics simulations (AIMD) of aqueous fluids were performed for fixed temperature and volume using the Car-Parrinello scheme \citep{CPMD1}. The temperature during each run was controlled by a Nos{\'e}--Hoover chain thermostat \citep{NK83,H85}. For each $T-V$ conditions at least $10\rm\,ps$ long trajectories were generated with an integration step of $0.12\rm\, fs$. \begin{figure*}[t] \includegraphics[angle=270,width=3.5in]{fig3.eps} \caption{Left panel: $\beta$ factors for $\rm H_3BO_3$ (solid line) and $\rm H_4BO_4^-$ (dashed lines) in aqueous solution. Right panel: fractionation factors between $\rm H_3BO_3$ and $\rm H_4BO_4^-$ in aqueous solution. The thick lines represent our results, while the thin lines represent the results of \citet{SV05} obtained using harmonic frequencies (their Table 2). The dotted line represents the corrected \citet{SV05} results. The correction is made by comparing the work of \citet{RB10} and the correction derivation procedure is discussed in the text. \label{F5}} \end{figure*} \subsection{Computational technique\label{CA}} The calculations of pseudofrequencies and $\beta$ factors for solids and aqueous solutions were performed by applying density functional theory (DFT) methods, which are currently the most efficient methods allowing for treating extended many particle systems quantum-mechanically. We used the planewave DFT code CPMD \citep{CPMD2}, which is especially suited for {\it ab initio} simulations of fluids, the BLYP exchange-correlation functional \citep{BECKE88,LEE88} and norm-conserving Goedecker pseudopotentials for the description of the core electrons \citep{GOE96}. One advantage of using the BLYP functional is that it usually gives harmonic frequencies that most closely resemble the observed frequencies of benchmark chemical systems \citep{FS95,AZZ10} \footnote{If the computed harmonic frequencies are closer to the observed, anharmonic frequencies than to the real harmonic frequencies of the computed system the resulted fractionation factor $\beta$ (Eq. \ref{beta}) computed from these frequencies should be closer to the real fractionation factor than the $\beta$ computed on a set of accurate harmonic frequencies.}. The energy cut-off for the plane wave basis set was $70\,\rm Ryd$ for geometry relaxations and molecular dynamics simulations and $140\rm\,Ryd$ for computation of vibrational frequencies. Periodic boundary conditions were applied for both crystalline solids and aqueous solutions to preserve the continuity of the media. \begin{table*}[t] \caption{The $1000(\beta-1)$, $\alpha=\beta_{\rm H_3BO_3}/\beta_{\rm H_4BO_4^-}$ and $\Delta=1000(\ln\beta_{\rm H_3BO_3}-\ln\beta_{\rm H_4BO_4^-})$ factors computed for isolated $\rm H_3BO_3$ and $\rm H_4BO_4^-$ at $T\rm=300\,K$ by \citet{Z05} using different basis sets and our result obtained using the full normal mode spectrum and equation \ref{beta}. The units are $\permil$.} \label{T3} \centering \begin{tabular}{lcccc} \hline\hline & $\rm H_3BO_3$ & $\rm H_4BO_4^-$ & $\alpha=\beta_{\rm H_3BO_3}/\beta_{\rm H_4BO_4^-}$ & $\Delta=1000(\ln\beta_{\rm H_3BO_3}-\ln\beta_{\rm H_4BO_4^-})$ \\ \hline 6-31+G(d) & 216.6 & 174.4 & 1.0359 & 35.3 \\ 6-311+G(d,p) & 215.0 & 170.5 & 1.0380 & 37.3 \\ our work & 211.3 & 167.3 & 1.0377 & 37.0 \\ \hline \end{tabular} \\ \end{table*} The force constants and frequencies needed for the computation of the $\beta$ factors were computed using the finite displacement scheme. Before performing the calculations of the crystal structures all atomic positions were relaxed to the equilibrium positions to minimize the forces acting on the atoms. We note that to compute the $\beta$ factors for crystals one formally should account for phonon dispersion. Here we use large supercells and restrict our calculations to a single phonon wave-vector ($\Gamma$). \citet{SCh11} has shown recently for $\rm ^{26}Mg/^{24}Mg$ fractionation in Mg-bearing minerals that supercells containing more than 20 atoms are sufficient to get very accurate $\beta$ factors even at $T=\rm 300\,K$ (error of 0.1\permil). At $T=1000\rm\,K$ the error is in the order of 0.01\permil. The accuracy of the high temperature isotope fractionation factors computed on a single phonon wave-vector is also demonstrated for iron-bearing minerals by \citet{BP09} and confirmed with good agreement of the predicted with the measured Li isotope fractionation factors between staurolite, spodumene, micas and aqueous fluid presented in our previous work \citep{KJ11}. Prior to the computation of the force constants and frequencies of boron atoms in the fluids the positions of all the atoms constituting the boron-carrying molecule ($\rm H_3BO_3$ or $\rm H_4BO_4^-$) were relaxed to the equilibrium positions, while all other atomic positions remained unchanged. The full normal mode analyzes were performed using the same method, but displacing all the atoms constituting the considered system. In the latter case the frequencies were obtained through the diagonalization of the full dynamical matrix \citep{S04} as implemented in CPMD code. The effect of the various approximations on the derived fractionation factors was studied by additional computations of $\rm H_3BO_3$ and $\rm H_4BO_4^-$ isolated clusters. For that purpose we used a large, isolated simulation box with a cell length of $16\rm\,\AA$, forcing the charge density to be zero at the boundary, as implemented in CPMD code. In order to compute the $\beta$ factors of boron species in the aqueous fluid we apply the same method as in our recent work on Li isotopes \citep{KJ11}, with the exception that we use the pseudofrequencies, i.e. the frequencies obtained from the three force constants acting on the fractionating element, and formula \ref{beta} for calculation of $\beta$ factors, as discussed in section \ref{SAA}. In order to fully account for the spatial continuity of the fluid and its dynamical motion we produced $10\rm\,ps$ long molecular dynamics trajectories of systems consisting of 64 H$_2$O molecules and one $\rm H_3BO_3$ or $\rm H_4BO_4^-$ molecule for different $T=1000\rm\,K$, $800\rm\,K$ and $600\rm\,K$ and pressure of $0.5\rm\,GPa$, which closely resembles the experimental conditions of \citet{WM05} and \citet{MW08}. The corresponding simulation box length is $13.75\rm\,\AA$ at $T=1000\rm\,K$. The $\beta$ factors were computed on the ionic configuration snapshots extracted uniformly in $0.1\rm\,ps$ intervals along the molecular dynamics trajectories. \subsection{Error estimation technique\label{ET}} The errors in the computed value of the $(\beta-1)$ and $\Delta$ fractionation factors were estimated from an average error of vibrational frequencies computed using the chosen DFT method. \citet{FS95}, \citet{MT02} and \citet{AZZ10} estimated the errors made in calculations of vibrational frequencies of small molecules using different DFT functionals. According to these works the BLYP functional systematically overestimates the harmonic frequencies by $\sim3.5\,\%$, with a deviation from the mean offset of $\sim1\,\%$. Therefore, we expect that using BLYP functional the $(\beta-1)$ and $\Delta$ values are systematically overestimated by $\rm 7\,\%$ and that in addition there is a $\rm 2\,\%$ error in derived $(\beta-1)$ factors. Similar errors result from using other functionals or even more sophisticated and time consuming post-Hartree-Fock methods such as MP2 \citep{FS95,AZZ10}. \section{Results and discussion} \subsection{Test of the computational method \label{TEST}} First, we illustrate the performance of the approximation proposed in section \ref{SAA} by computing the $\beta$ factors for the isolated $\rm H_3BO_3$ and $\rm H_4BO_4^-$ molecules and selected crystalline solids. In Figure \ref{F1} we present three sets of calculations of $\beta$ factors: (1) the ``exact" result obtained from a full normal mode analysis and formula \ref{beta}, (2) the results obtained applying \citet{KJ11} method based on Eq. \ref{BAPX}, (3) the results obtained using pseudofrequencies computed for the fractionating element and Eq. \ref{beta} for the estimation of the $\beta$ factor. The numerical values for selected temperatures are reported in Table \ref{T1}. Approach (3) results in much better agreement with the ``exact" result. For $\rm H_3BO_3$, the $\beta$ factor is overestimated by only $0.5\permil$ and $1.5\permil$ for temperatures of $\rm 800\,K$ and $\rm 600\,K$ respectively. Applying method (2), the error is more pronounced, $2.2\permil$ and $6.7\permil$ respectively. In the case of molecular $\rm H_4BO_4^-$, the errors using method (3) for the same temperatures are only $0.1\permil$ and $0.5\permil$ respectively. The same behavior is shown for dravite and boromuscovite crystalline solids that contain boron in the coordinative arrangement that resemble the configurations of aforementioned B-bearing molecules. For $T>600\rm \,K$ the proposed method represents only a few percent correction to the approximation given by equation (\ref{BAPX}), so the relation (\ref{TESTP}) is satisfied. It is evident that for B-bearing materials considered here the improvement made by using the pseudofrequencies based approach is substantial. It corrects for about 75$\%$ of error of the \citet{BM47} approximation (Eq. \ref{BAPX}). However, the question of general applicability of the proposed method to other isotopic systems would require careful testing on a large set of materials, which is well beyond the scope of the current paper. \subsection{B isotope fractionation in gas and fluid phases} \subsubsection{B isotope fractionation between $\rm H_3BO_3$ and $\rm H_4BO_4^-$ in the gas phase} In a first step of our investigation of boron-rich aqueous fluids we derived the full frequency spectra of molecules in the gas phase (isolated molecules). The relevant $\beta$ factors were computed using Equation \ref{beta}. These studies were performed in order to compare our results with the published values of \citet{Z05}, both computed using the same DFT BLYP functional. In Table \ref{T1a} we report the computed frequencies that are affected by the different B isotope substitutions along with other theoretical estimations and experimental measurements. The computed frequencies are in good agreement with earlier theoretical predictions and show similar agreement with the experimental measurements. The results in terms of computed $\beta$ factors for the two considered species are reported in Figure \ref{F6}, where we compare our results with the values computed using frequencies of \citet{Z05}. The comparison of the two sets of calculations reveals that our $\beta$ factors for both species are smaller by $\sim1\,\permil$ at $\rm 600\, K-1000\,K$ than the values of \citet{Z05}. However, the difference between $\beta$ factors of $\rm H_3BO_3$ and $\rm H_4BO_4^-$ remains nearly identical in both sets of calculations and the agreement is nearly perfect for higher temperatures. We note, that for the comparison we used the frequencies of \citet{Z05} computed using 6-31+G(d) basis set, as only these are provided by the authors. $\beta$ and $\alpha$ factors obtained at $T=300\rm\,K$ using a more extended 6-311+G(d,p) basis set indicate that the $\beta$ factors using 6-31+G(d) basis set are not fully converged. In particular, the $(\beta-1)$ factor of $\rm H_4BO_4^-$ computed with 6-311+G(d,p) basis set is $\rm 3.9\,\permil$ smaller than the one derived using 6-31+G(d). In Table \ref{T3}, we compare these results with the results of our calculation. It is clearly seen that for lower temperatures such as $T=300\rm\,K$ the values computed with 6-311+G(d,p) basis set are in better agreement with our results indicating that plane-wave based DFT approach we use provides adequate vibrational frequencies and resulting isotope fractionation factors. \begin{figure*}[t] \includegraphics[angle=270,width=3.5in]{fig4.eps} \caption{Left panel: pressure dependence of the $\beta$ factor of neutral fluid ($\rm H_3BO_3$) at $T\rm=1000\,K$. Dashed line is the linear regression fit to the calculated values (points): $1000(\beta-1)=23.6+0.28\,P{\rm (GPa)}$; Right panel: The computed change of the $\beta$ factor with pressure (symbols) in comparison to the increase in the $\beta$ factor derived from the frequency shifts of the $666\,\rm cm^{-1}$ and $1454\,\rm cm^{-1}$ lines measured by \citet{SV05} (dashed line). The comparison is made assuming that $(\beta-1)\sim\nu^2$. \label{F7a}} \end{figure*} \begin{figure*}[t] \includegraphics[angle=270,width=3.5in]{fig5.eps} \caption{Left panel: pressure dependence of the $\beta$ factor of strongly basic fluid ($\rm H_4BO_4^-$) at $T\rm=1000\,K$. Dashed line is the regression fit to the calculated values (points): $1000(\beta-1)=17.15+0.754\,P-0.027\,P^2$, where pressure is given in $\rm GPa$; Right panel: The computed change of the $\beta$ factor with pressure in comparison to the increase in the $\beta$ factor derived from the frequency shift of the $975\,\rm cm^{-1}$ line measured by \citet{SV05} (dashed line). The comparison is made assuming that $(\beta-1)\sim\nu^2$. \label{F7b}} \end{figure*} \subsubsection{B isotope fractionation between $\rm H_3BO_3$ and $\rm H_4BO_4^-$ in aqueous fluid} In order to obtain the temperature dependent $\beta$ factor for aqueous fluids we fitted the function $1+A/T^2+B/T^4$ to the computed values using the least squares minimization procedure. The computed $\beta$ values for $\rm H_3BO_3$ in fluid are: $1.02366\pm0.00012$, $1.03624\pm0.00018$ and $1.06262\pm0.00010$ and for $\rm H_4BO_4^-$ in fluid are: $1.01745\pm0.00005$, $1.02650\pm0.00010$ and $1.04597\pm0.00015$, for the temperatures of $1000\rm \,K$, $800\rm \,K$ and $600\rm \,K$ respectively. The resulting temperature dependent $\beta$ factor for $\rm H_3BO_3$ is $\beta=1+2.416\cdot 10^4/T^2-5.823\cdot 10^{8}/T^4$ and for $\rm H_4BO_4^-$ is $\beta=1+1.772\cdot 10^4/T^2-4.234\cdot 10^{8}/T^4$. The results for $\rm H_3BO_3$ and $\rm H_4BO_4^-$ in aqueous solutions are shown in Figure \ref{F5}. As was observed for the isolated molecules, the $\beta$ factor of $\rm H_3BO_3$-bearing fluid is substantially larger than the one for the $\rm H_4BO_4^-$. This can be understood in terms of the substantial difference in the B-O bond lengths exhibited by the two considered species. In case of isolated molecules our calculations indicate a B-O bond length of $1.40\rm\,\AA$ for $\rm H_3BO_3$ and $1.51\rm\,\AA$ for $\rm H_4BO_4^-$. We compared our $\beta$ factors with the values computed by \citet{SV05}, which were derived by the combination of force field methods and experimental data to derive accurate vibrational frequencies. For $\rm H_3BO_3$ we got a nearly identical result. In case of $\rm H_4BO_4^-$ our calculation predicts a value which is lower by $\rm 2-4\,\permil$. However, \citet{RB10} and \citet{RB07} revealed the improper assignment of a major fractionating vibrational mode of $\rm H_4BO_4^-$ in the force field by \citet{SV05}. This leads to the underestimation of the fractionation factor between aqueous $\rm H_3BO_3$ and $\rm H_4BO_4^-$ by \citet{SV05}. Assuming that $\alpha\propto T^{-2}$ and having the difference between BLYP calculations of \citet{RB10} and \citet{SV05} of $\Delta\alpha=16.4\,\permil$ at $T\rm=300\,K$, the value reported by \citet{SV05} should be underestimated by $\Delta\alpha=16.4\cdot 2(300/T)^2\,\permil$, which results in $\Delta\alpha\sim1.5\,\permil$ at $T\rm=1000\,K$. Corrected in such as way result of \citet{SV05} is also plotted in Figure \ref{F5}. It is now very consistent with our prediction. \subsubsection{Discussion of computational errors} Most previous computational studies of boron isotope fractionation in aqueous solutions concentrate on the computation of the isotope fractionation at ambient conditions \citep{RB10,RB07,LT05,Z05}. \citet{RB10} performed detailed analysis of impact of the chosen computational method (HF, MP2, different DFT functionals) and size of the basis set on the calculated fractionation factors between $\rm H_3BO_3$ and $\rm H_4BO_4^-$. They found that DFT methods are not performing well for the borate system and concluded that DFT {\it ''is of limited usefulness in chemically accurate predictions of isotope fractionation in aqueous systems"} \citep{RB10}. The empirically derived error of the derived fractionation factor is of the order of $5-10\,\permil$ for a total fractionation of $\sim30\rm \,\permil$. We note that this is expected and clearly visible if we apply the error estimation procedure outlined in section \ref{ET}. For instance, at room temperature the derived beta factors using the BLYP functional are $213.6\,\permil$ and $173.3\,\permil$ respectively (\citet{RB10}, Table 2). This gives a fractionation factor of 1.0343. Following our error estimation scheme, the absolute error of the fractionation factor is $10.5\,\permil$, and the properly reported computed value is $\alpha=1.034\pm 0.011$. When one corrects for the systematic error of $7\%$ and assumes $2\%$ of statistical error on $\beta$ factors, then the value of $\alpha$ decreases and the error is slightly smaller, i.e. $\alpha=1.032\pm 0.008$. This is in good agreement with the experimental data reported in \citet{RB10} and explains the spread of the values computed using different methods and reported in that paper. It is very difficult to get the fractionation factors for ambient conditions, as the fractionation factor is often just a small fraction of the relevant $(\beta-1)$ factors, $(\alpha-1)\sim0.15(\beta-1)$ in the considered case. Assuming that $(\alpha-1)=0.15(\beta-1)$, a $2\%$ error in the $(\beta-1)$ factors leads to an absolute error in $(\alpha-1)$ of $0.04(\beta-1)=0.04(\alpha-1)/0.15\sim0.27(\alpha-1)$, i.e. $\sim 27\%$ of relative error in the derived fractionation factor $(\alpha-1)$. On the other hand, we note that such a big error is not substantially larger than the uncertainties in the experimental data reported by \citet{RB10} in their Figure 2. Thus, the case of boron fractionation in aqueous fluid at ambient conditions does not necessarily show the limited usefulness of DFT in the prediction of isotope fractionation factors, but only reflects the fact that precise estimation or measurement of the B isotopes fractionation factors at ambient condition requires unprecedented accuracy of both experimental or computational techniques. For instance, in order to get the value of $(\alpha-1)$ with a relative error of $5\%$ (at ambient conditions) one needs to estimate the $(\beta-1)$ factors or measure relevant quantities with precision of less than $1\%$. At higher temperatures the situation is different. Looking just at the fractionation between $\rm H_3BO_3$ and $\rm H_4BO_4^-$ in the gas phase or the aqueous solution one can see that for $T>600\rm\,K$ the fractionation factor between the two substances, $(\alpha-1)$, is at least $25\%$ of the $(\beta-1)$ factor. This results in smaller $0.04/0.25\sim 16\%$ for $T=600\rm\,K$ and $0.04/0.36\sim 11\%$ for $T=1000\rm\,K$ relative error, which is acceptable in our calculations. Nevertheless, this case shows the importance of proper error estimation on the computed fractionation factors. Such an estimation is usually omitted or not provided explicitly, which can lead to wrong conclusions when the theoretical prediction is confronted with the measured data. \subsubsection{Pressure dependence of the fluid fractionation factor} In our recent paper \citep{KJ11} we have shown that due to compression the $\beta$ factor of Li in aqueous fluid increases with increase in pressure (for $P\rm>2\rm \,GPa$). The same should happen for $\rm H_3BO_3$ and $\rm H_4BO_4^-$ aqueous fluid as the vibrational frequencies of boron species in aqueous fluid increase with increase in pressure \citep{SV05,STH05}. Having the experimental data we checked whether the derived pressure-dependent $\beta$ factors are consistent with the pressure shifts of vibrational frequencies of considered boron species measured by \citet{SV05}. For that purpose we performed a set of calculations using supercells containing 8 water molecules and the relevant boron species. We note that in line with our previous results for Li \citep{KJ11}, the obtained values of ($\beta-1$) at $P\rm=0.5\,GPa$ are within $\rm 0.1\,\permil$ in agreement with the values obtained for supercells containing 64 water molecules. The results are given in Figures \ref{F7a} and \ref{F7b}. The computed $(\beta-1)$ values for $\rm H_3BO_3$ fluid show a linear dependence in pressure, $(\beta-1)=23.60+0.28\,P\rm(GPa)\, \permil$. This is expected as $\rm (\beta-1)\propto\nu^2\sim\nu_0^2+2\nu_0\Delta\nu$ \citep{S04} and $\Delta\nu$ is a linear function of pressure \citep{SV05,STH05}. In case of $\rm H_4BO_4^-$ the pressure-dependence is linear up to $P\rm \sim2-3\,GPa$ and it becomes less steep at higher pressures. In order to quantitatively check the consistency of our prediction with the measured vibrational frequency shifts of \citet{SV05} we derived the relative shifts in the $(\beta-1)$ factor assuming that $\rm (\beta-1)\propto\nu^2$ and the measured pressure dependence of the frequency shifts: $\Delta \nu=2.15{\rm\,cm^{-1}}\cdot P(\rm GPa)$ and $\Delta \nu=3.50{\rm\,cm^{-1}}\cdot P(\rm GPa)$ for $\rm 1454\,cm^{-1}$ and $\rm 666\,cm^{-1}$ vibrational frequencies of $\rm H_3BO_3$ and $\Delta \nu=6.47{\rm\,cm^{-1}}\cdot P(\rm GPa)$ for the $\rm 975\,cm^{-1}$ vibrational frequency of $\rm H_4BO_4^-$. The chosen vibrational frequencies are the ones affected by the different B isotope substitution. Our predicted shift of $(\beta-1)$ matches well the shifts derived from the measured frequency shifts. Such a good agreement with the experimental data validates further our computational approach and shows that {\it ab initio} calculations can be successfully used in derivation of the pressure dependence of the fractionation factors and pressure-induced vibrational frequencies shifts. Moreover, first principles calculations can be useful in extrapolation of the experimental values for $\beta$ and $\Delta\nu$ to more extreme conditions, which otherwise are extremely difficult to reach by experimental techniques. \subsection{Fluid-mineral fractionation} Next we present the results of the fractionation between boron bearing fluids and minerals such as dravite, olenite and boromuscovite. The aim of these studies is to investigate the mechanisms driving the fractionation process, the role of coordination and the B-O bond length. Below we discuss each case separately. \begin{figure}[t] \includegraphics[angle=270,width=2.0in]{fig6.eps} \caption{The fractionation factors between tourmaline and aqueous fluid. The solid line represents our prediction for the fractionation between dravite and $\rm H_3BO_3$ neutral fluid and the shadowed region represent the computational uncertainty. The dotted line represents our prediction for the fractionation between olenite containing $\rm B^{[3]}$ species only and neutral fluid. The computational error is similar in both cases. The data points are the values measured for dravite-fluid system by \citet{MW08}. \label{F3}} \end{figure} \subsubsection{Tourmaline-neutral fluid} \citet{MW08} measured the boron isotope fractionation between tourmaline and neutral fluid at $T=400-700\,\rm ^{o}C$ and $P\rm=0.2\,GPa$. In the experiment the tourmaline was represented by dravite. In contrast to the former measurements of \citet{P92} the measured fractionation is very small and does not exceed $2.5\,\permil$ at $400\rm\,^{o}C$. Our calculated fractionation curve together with the experimental data are given in Figure \ref{F3}. Our result correctly reproduces the experimental measurements within the computational accuracy. The dravite-fluid fractionation is small as the two materials contain boron in $\rm BO_3$ units. We also predict a small fractionation between olenite carrying 3-fold coordinated boron only and aqueous fluid, although the olenite-fluid fractionation is positive because of the shorter B-O bond lengths for olenite ($\rm 1.378\,\AA$ vs. $\rm 1.397\,\AA$). \begin{figure}[t] \includegraphics[angle=270,width=2.0in]{fig7.eps} \caption{The fractionation factors between boromuscovite and basic aqueous fluid (fluid containing $\rm H_4BO_4^-$). The lines represent our results assuming the presence of boron species in form of $\rm H_4BO_4^-$ only (solid line) and mixture of $90\%$ of $\rm H_4BO_4^-$ and $10\%$ of $\rm H_3BO_3$ (dotted line) in the fluid. The data points are the values measured by \citet{WM05}. The uncertainty of calculated values is indicated by shadowed area and is similar for both curves. \label{F4}} \end{figure} \subsubsection{Boromuscovite-strongly basic fluid \label{SB}} Boromuscovite synthesized in the experiments of \citet{WM05} consisted of two type of polytypes, 1M ($\sim10\,\%$) and 2M1 ($\sim90\,\%$). In order to be consistent with the experimental conditions, we derived the $\beta$ factors for both polytypes and computed their weighted average. We note that the $\beta$ factors for both polytypes of mica are similar with the difference in $(\beta-1)$ not larger than $3\rm\,\%$. This is consistent with similar B-O bond lengths ($1.532\rm\,\AA$) found in both polytypes. Boromuscovite contains boron in tetrahedral sites. Therefore, in order to investigate the impact of the B-O bond length on the fractionation we first compare the fractionation between the mineral and a strongly basic fluid containing boron in $\rm H_4BO_4^-$. The result, together with the measurements of \citet{WM05} of the fractionation between boromuscovite and strongly basic fluid, is summarized in Figure \ref{F4}. Our calculations predict a negative fractionation between mica and the $\rm H_4BO_4^-$ fluid. The agreement of our prediction with the experimental measurements is relatively good; however, the experimental data indicate slightly stronger fractionation. We note that the experimental conditions of \citet{WM05} do not assure that the measured basic fluid contained four-fold coordinated boron species, i.e. $\rm H_4BO_4^-$, only. As we indicate in the Figure \ref{F4}, the presence of as little as $10\%$ of $\rm H_3BO_3$ in the measured basic fluid brings the prediction and measurements into much better agreement. It makes sense that the $\beta$ factor of boromuscovite is smaller than for aqueous $\rm H_4BO_4^-$ as the average B-O bond length in mica is $\rm 1.532\,\AA$, while it is $\rm 1.513\,\AA$ and therefore shorter in case of aqueous $\rm H_4BO_4^-$. We note that in our derivation we assumed that the fluid consists mostly of $^{[4]}\rm B$ species. As we already mentioned, although \citet{WM05} call the fluid ``strongly basic" its exact composition, especially the amount of $^{[3]}\rm B$ species is unknown. However, if for instance the $^{[4]}\rm B$ to $^{[3]}\rm B$ ratio was $1$, then the predicted mica-basic fluid fractionation at $\rm 800\,K$ would be $5\,\permil$ larger than measured. This would result in a large inconsistency between the computed values and the experimental data. On the other hand, good agreement between the prediction and the measurements indicates that the strongly basic fluid was dominated by $\rm H_4BO_4^-$ species, which is in line with previous studies \citep{Z05,SV05}. \begin{figure}[t] \includegraphics[angle=270,width=2.0in]{fig8.eps} \caption{The fractionation factors between boromuscovite and aqueous fluid. The data points are the values measured by \citet{WM05}. The solid line represents the result obtained for ambient pressure. The dashed line represents the result for fluid containing $\rm H_3BO_3$ only obtained for $P=3\,\rm GPa$ accounting for compression and thermal expansion, with the uncertainties in calculated values indicated by shadowed area. The dotted lines represent the results assuming different admixture of four-fold coordinated boron species (represented by $\rm H_4BO_4^-$ with abundance indicated in the figure) to the fluid. The computational error is comparable for all the results. \label{F8}} \end{figure} \subsubsection{Boromuscovite-neutral fluid \label{BNF}} The fractionation between boromuscovite and neutral fluid involves a change in coordination from $\rm B^{[4]}$ in boromuscovite to $\rm B^{[3]}$ in neutral fluid. \citet{WM05} measured the fractionation between the two materials at $\rm 3\,GPa$, shown in Figure \ref{F8}. The predicted fractionation is about $3\pm2.5\,\permil$ larger than the measured value. Looking for potential sources of this discrepancy, we have checked for the effect of the change in lattice parameters due to combined thermal expansion and compression. For that purpose we applied the EOS of \citet{HP11} for muscovite, which gives $4.4\%$, $3.8\%$ and $3.1\%$ decrease in volume for $T=600\,\rm K$, $800\,\rm K$ and $1000\,\rm K$ respectively and $P=3\rm\,GPa$. As muscovites show highly anisotropic compressibility patterns, in line with \citet{CZ95} we applied the $T$ and $P$ driven change in volume assuming the $16\%$, $19\%$ and $65\%$ contribution to compression along the $\vec{a}$, $\vec{b}$, $\vec{c}$ lattice vectors. On the other hand $(\beta-1)$ factors of $\rm H_3BO_3$ and $\rm H_4BO_4^-$ in aqueous solutions at $P\rm=3\,GPa$ increase by $3.5\%$ and $11.8\%$ respectively (Figures \ref{F7a} and \ref{F7b}), leading to pressure-induced increase in the boromuscovite-aqueous fluid fractionation at the experimental pressure. The $\Delta$ factor, corrected for effects of thermal expansion and compression of boromuscovite and compression of fluid, is also given in Figure \ref{F8}. Because the high $T$ and $P$ effects result in similar increases in the $\beta$ factors for both solid and aqueous fluid, the resulting fractionation factor between these two phases is close to the one derived at ambient conditions. Therefore, thermal expansion and compression effects cannot explain the observed discrepancy between prediction and the measurements of \citet{WM05}. On the other hand, the comparison of our results with the experimental data suggests that the fractionation between boromuscovite and fluid is the same as between $\rm H_3BO_3$ and $\rm H_4BO_4^-$ fluids (see Figure \ref{F5}), which is at odds with the non-negligible and negative fractionation between boromuscovite and a strongly basic fluid. In section \ref{SB} we have shown that we are able to correctly reproduce the fractionation between boromuscovite and strongly basic fluid, which indicates that our result for boromuscovite is reliable. This suggests that another, unaccounted effect leads to the decrease of the boron isotope fractionation between mica and neutral fluid in the experiments of \citet{WM05}. One possible solution for the discrepancy is a non-negligible amount of boron residing in four-fold coordinated configurations in neutral solution. This is in line with the Raman spectroscopy measurements of \citet{STH05}, who detected a broad peak in the Raman spectra of neutral $\rm H_3BO_3$-dominated fluid and attributed it to $\rm B^{[4]}$ species. The integrated area of this peak, compared to the peak of the Raman $877\,\rm cm^{-1}$ line of $\rm B^{[3]}$ species, indicates the presence of at least $15-30\,\%$ of $\rm B^{[4]}$ species by mole fraction. Assuming that there is $\rm 15-30\,\%$ of $\rm B^{[4]}$ species present in the fluid and that the $\beta$ factor of these species is similar to that of $\rm H_4BO_4^-$, the fractionation factor between boromuscovite and $\rm H_3BO_3$ aqueous fluid decreases bringing the theory and the experiment to better agreement, which is illustrated in Figure \ref{F8}. If this interpretation is true, it suggests that boron isotope fractionation could be used to gather information on the speciation of B in aqueous fluids. \begin{figure}[t] \includegraphics[angle=270,width=2.0in]{fig9.eps} \caption{ The fractionation factors between mica and tourmaline. The solid line represents our value for fractionation between B-muscovite and dravite. The dashed line is the experimental fractionation factor between tourmaline and mica determined by \citet{WM05} and \citet{MW08}. The experimental error is $\rm 2\,\permil$. The dotted line is the experimental fractionation factor of \citet{WM05} and \citet{MW08} but corrected for the presence of $\rm B^{[4]}$ species in the neutral fluid in the high $P$ experiments of \citet{WM05}, as is discussed in the text. The diamonds are the data from natural samples taken from \citet{KM11} and references herein. The uncertainties in calculated values are indicated by shadowed area. \label{F2}} \end{figure} \subsection{B isotope fractionation between minerals} The boron isotope fractionation between B-bearing crystalline solids has received considerable attention recently \citep{WM05,MW08,KM11,M05,H02}. We focus here on the investigation of boron isotope fractionation between mica and tourmaline as boron atoms in these minerals occupy sites of different coordination, which should result in a large B isotope fractionation between these two minerals. In micas boron substitutes for silicon in the four-fold coordinated site \citep{WM05}, while in tourmaline (dravite) it is incorporated in the three fold coordinated site \citep{MW08}. Comparing the B isotope fractionation between different minerals, melts and fluids \citet{WM05} have shown that the fractionation between two materials of different B coordination is large, reaching $5\,\permil$ at $\rm 1000\,K$ and much higher values at lower temperatures. \citet{KM11},\citet{M05} and \citet{H02} measured the fractionation between coexisting phases of the two minerals in natural samples. The fractionation between these two minerals is also derived from experimental isotopic fractionation data of B-muscovite-fluid \citep{WM05} and tourmaline-fluid \citep{MW08} systems. The results of these measurements and our computed $T$-dependent fractionation curve are given in Figure \ref{F2}. The first striking observation is that our predicted fractionation factors are much larger (taking the absolute value) than the experimental values \citep{WM05,MW08}. The latter are also inconsistent with the natural samples data of \citet{KM11} and previous studies discussed in that paper \citep{M05,H02}. On the other hand, the measurements on natural samples are consistent with our calculated values, which tends to validate our predictions. We notice that the most recent measurements of boron isotope signatures of tourmaline and white mica from the Broken Hill area in Australia by \citet{KM11} indicate for the assumed temperature of 600 $^{\rm o}$C that the fractionation factor between the two phases is 10.4$\pm2.7\,\permil$, which is in good agreement with our computed value of $10.7\pm1.8 \, \permil$. The experimental mica--tourmaline B isotope fractionation factors of \citet{WM05} and \citet{MW08} are $2\,\rm\permil$ and $6\,\rm\permil$ smaller at temperatures of $1000\,\rm K$ and $800\,\rm K$ respectively, with an experimental uncertainty of $2\rm\,\permil$. However, this discrepancy can be resolved by assuming that in the experiments of \citet{WM05} the fluid contained a significant admixture of $\rm B^{\rm [4]}$ species, which leads to the underestimation of the experimental boromuscovite-fluid fractionation factor by $\sim2\,\permil$ at $1000\,\rm K$ and $\sim3.5\,\permil$ at $1000\,\rm K$, as is seen in Figure \ref{F8}. The experimental mica--tourmaline fractionation factor corrected for the presence of $\rm B^{\rm [4]}$ species is also plotted in Figure \ref{F2}. It is now more consistent with the natural data. We note that this result independently supports the conclusion underlined in section \ref{BNF} and result of \citet{STH05} that high-$P$, B-bearing neutral fluids contain significant admixtures of $\rm B^{\rm [4]}$ species. Olenite is a mineral which can incorporate boron in both trigonal and tetrahedral sites as it substitutes for both Al and Si atoms. It is therefore interesting to check the fractionation of boron isotopes between the two differently coordinated sites in one mineral and compare it with the above result for mica and tourmaline. The computed fractionation between the trigonal and tetrahedral sites at 600 $^{\rm o}$C is $10.6\pm1.9\,\permil$, which is consistent with the fractionation between tourmaline and mica, indicating that the coordination of the B atom is the driving factor for the fractionation of the B isotopes. Similarly, we computed the boron isotopes fractionation between trigonal and tetragonal boron sites in dravite. In order to create the tetragonal B site we replaced one Si atom with B and we added one H atom forming an additional OH group to compensate the charge. The computed fractionation between the sites at 600 $^{\rm o}$C is $8.9\pm1.7\,\permil$, which is also in agreement with the aforementioned results. Next, we will show that the value of the $\beta$ factor depends not only on coordination but is also strongly correlated with B-O bond length. \begin{figure}[t] \includegraphics[angle=270,width=2.0in]{fig10.eps} \caption{$\beta$ factors for various considered materials as a function of temperature. The materials are indicated on the right side. Their order reflects the value of the $\beta$ factor at $1000\,\rm K$ from the largest (top) to the smallest (bottom). \label{F10}} \end{figure} \subsection{ Fractionation between $\rm B^{[3]}$ and $\rm B^{[4]}$ materials} The $\beta$ factors computed for all the considered materials are grouped together in Figure \ref{F10}. $\beta$ factors can be grouped into two sets, one that includes materials with boron in three-fold coordination and another one that includes materials having boron in four-fold coordination. For olenite and dravite we also computed the $\beta$ factors with boron sitting on four-fold coordinated site. The $\beta$ factor for these crystalline solids with given $\rm B^{[3]}/B^{[4]}$ ratio can be derived as a weighted average of the $\beta$ factors obtained for boron sitting on the two differently coordinated sites. The fractionation factor between materials of different boron coordination is $\sim8\,\permil$ on average at $T=1000\,\rm K$. We note that it is $\sim3\,\permil$ larger than the one deduced by \citet{WM05} from measurements performed on solids, silicate melts and fluids, but this can be attributed to the underestimation of the fractionation factors for boromuscovite-fluid system by \citet{WM05} due to potential admixture of $\rm B^{[4]}$ species in the investigated fluid. \begin{figure}[t] \includegraphics[angle=270,width=2.0in]{fig11.eps} \caption{The $\beta$ factor at $T\rm=1000\,K$ for various considered materials as a function of B-O bond length. Filled circles represent the values obtained for boron in trigonal sites and open circles represent the values obtained for tetragonal sites. \label{F7}} \end{figure} Our results show also a substantial spread of $\beta$ factors of substances containing boron of a given coordination. The spread is at least $4\,\permil$ and results from different B-O bond lengths. We illustrate this in Figure \ref{F7} by plotting together the $\beta$ factors derived for all considered materials at $T=1000\rm\,K$ as a function of B-O bond length. It is clearly seen that there is a roughly linear correlation between the $\beta$ factor and B-O bond length, which is especially evident comparing the results for crystalline solids. For instance, out of the considered $\rm B^{[4]}$-bearing minerals boromuscovite has the longest B-O bond length of $\rm 1.525\,\AA$ (1M) and $\rm 1.516\,\AA$ (2M1), followed by dravite $\rm 1.514\,\AA$ and olenite with B-O bond length of $\rm 1.502\,\AA$. This tracks the differences in the $\beta$ factors derived for these materials. In addition the materials having B$^{[3]}$ species only exhibit shorter bond lengths of $\rm \sim1.37\,\AA$ and higher $\beta$ factors, while the materials containing B$^{[4]}$ species having bond lengths of about $\rm \sim1.52\,\AA$ show much smaller $\beta$ factors. Therefore, the tighter bonding of B$^{[3]}$ species likely explains why the heavy B isotope prefers the less coordinated phases. This clearly shows that the change in the B-O bond length during an isotope exchange is the leading factor driving the production of the boron equilibrium isotope signatures at high $T$. \section{Conclusions} In this work we have presented a detailed analysis of boron isotope fractionation between boron-bearing crystalline solids and aqueous fluids at high $T$ and $P$ conditions. In order to perform our investigation we have applied and extended a computationally efficient approach for the computation of isotope fractionation factors for complex minerals and fluids at high temperatures and pressures presented by \citet{KJ11}. As an extension to the \citet{BM47} {\it ``single atom approximation"} method we demonstrated that using the pseudofrequencies derived from the force constants acting on the fractionating element together with the full formula for computation of the reduced partition function ratios results in significant improvement in the accuracy of the computed fractionation factors, which is essential when lower temperature materials and high vibrational frequency complexes are considered. In order to understand the fractionation between B-bearing crystalline solids and aqueous fluids we performed a set of calculations of $\beta$ factors for dravite, olenite, boromuscovite and aqueous solutions of $\rm H_3BO_3$ and $\rm H_4BO_4^-$. In agreement with the experimental findings we show that the fractionation strongly correlates with coordination through the change in the B-O bond length. The lower trigonal coordination $\rm BO_3$ arrangement results in higher $\rm ^{11}B/^{10}B$ (by $\rm \sim 8\,\permil $ at $T=\rm1000\,K$) than the tetrahedrally coordinated boron complexes, which exhibit $\rm \sim0.15\,\AA$ longer B-O bonds. The computed fractionation between minerals and fluids of the same coordination are in good agreement with experiments. However, we predict larger isotope fractionation between boromuscovite and $\rm H_3BO_3$ fluid (by at least a few $\permil$) than was measured {\it in situ} at high $P$ by \citet{WM05} and \citet{MW08}, but that is consistent with measurements on natural samples. We note that the presence of $\rm B^{[4]}$ in high-$P$ fluid could reconcile the {\it in situ} experimental results with our prediction and other measurements. This is expected from the experiments of \citet{STH05}, but requires further experimental confirmation. If true, this would open the possibility for using the isotope fractionation techniques as a tool to measure the speciation of boron in fluids and crystalline solids. We have also demonstrated that with our computational approach we are able to correctly predict the pressure-induced isotope fractionation for compressed aqueous fluids, which indicates the ability of {\it ab initio} methods to predict the isotopic signatures of highly compressed materials, even those that are difficult to investigate experimentally. Our study confirms that {\it ab initio} computer simulations are a useful tool not only for prediction but also understanding the equilibrium stable isotope fractionation processes between various phases, including aqueous solutions, at high pressures and temperatures. They can nicely complement experimental efforts, provide unique insight into the isotope fractionation process on the atomic scale and deliver data for conditions that are inaccessible by the current experimental techniques. \section*{Acknowledgements} The authors wish to acknowledge financial support in the framework of DFG project no. JA 1469/4-1. Part of the calculations were performed on the IBM BlueGene/P JUGENE of the John von Neumann Institute for Computing (NIC). We are also grateful the associate editor Edwin A. Schauble and anonymous referees for constructive comments that helped improving the manuscript. \bibliographystyle{model1-num-names}
\section{Introduction} \label{sec:intro} Wilson lines (also known as gauge links or eikonal lines) {can} be naturally introduced in any gauge field theory. These objects are generically defined via traces of path-ordered exponentials of a gauge field evaluated along a given trajectory $ {\cal W} (\Gamma) = {\cal P} \exp \left[\ -ig\int_{[{\Gamma}]}\ dz^{\mu} \ {\cal A}_{\mu}(z) \ \right] $. The path $\Gamma$ is a curve along which the gauge field ${\cal A}$ gets transported from the initial point to the final one. Wilson lines defined on closed contours are called Wilson loops. They are path-dependent non-local functionals of the gauge field, invariant under gauge group transformations. Putting the matter of question more mathematical, one can construct a space with its elements being Wilson loops defined on an infinite set of contours. Reformulation of QCD in terms of the elements of a generic loop space would allow one to use gauge-invariant quantities as fundamental degrees of freedom instead of the quarks and gluons from the standard QCD Lagrangian \cite{Loop_Space, WL_RG}. Observables can then be obtained via correlation functions of Wilson loops: \begin{equation} {\cal W}_n (\Gamma_1, ... \Gamma_n) = \Big \langle 0 \Big| {\cal T} \frac{1}{N_c} {\rm Tr}\ \Phi (\Gamma_1)\cdot \cdot \cdot \frac{1}{N_c}{\rm Tr}\ \Phi (\Gamma_n) \Big| 0 \Big\rangle \ , \ \Phi (\Gamma_i) = {\cal P} \ \exp\[ig \oint_{\Gamma_i} \ dz^\mu A_{\mu} (z) \] \ . \label{eq:wl_def} \end{equation} Complete information on the quantum dynamical properties of the loop space is accumulated in the Schwinger-Dyson equations: \begin{equation} \langle 0 | \nabla_\mu F^{\mu\nu} \ {\cal O} (A) \ | 0 \rangle = i \langle 0 | \frac{\delta }{\delta A_\nu} \ {\cal O} (A) \ | 0 \rangle \ , \label{eq:sch_dy_YM} \end{equation} where ${\cal O} (A)$ stands for an arbitrary functional of the gauge fields. Let the functionals ${\cal O} (A)$ be the Wilson exponentials $\Phi (\Gamma)$ (\ref{eq:wl_def}). Then Eqs. (\ref{eq:sch_dy_YM}) turn into the Makeenko-Migdal (MM) equations \cite{MM_WL}: \begin{equation} \partial_x^\nu \ \frac{\delta}{\delta \sigma_{\mu\nu} (x)} \ {\cal W}_1(\Gamma) = N_c g^2 \ \oint_{\Gamma} \ dz^\mu \ \delta^{(4)} (x - z) {\cal W}_2(\Gamma_{xz} \Gamma_{zx}) \ , \label{eq:MM_general} \end{equation} where the basic operations are the area- $\delta/\delta\sigma_{\mu\nu}$ and the path- $\partial_\mu$ derivatives \cite{MM_WL}: \begin{equation} \frac{\delta}{\delta \sigma_{\mu\nu} (x)} \ \Phi (\Gamma) \equiv \lim_{|\delta \sigma_{\mu\nu} (x)| \to 0} \ \frac{ \Phi (\Gamma\delta \Gamma) - \Phi (\Gamma) } {|\delta \sigma_{\mu\nu} (x)|} \ , \label{eq:area_derivative} \end{equation} and the contour $\Gamma\delta \Gamma$ is obtained from the initial one by means of the infinitesimal area deformation $\delta \Gamma$ at the point $x$, while the path variation without changing the area gives rise to the path derivative \begin{equation} \partial_\mu \Phi(\Gamma) = \lim_{|\delta x_{\mu}|} \frac{\Phi(\delta x_\mu^{-1}\Gamma\delta x_\mu) - \Phi(\Gamma)}{|\delta x_{\mu}|} \ . \label{eq:path_derivative} \end{equation} The area derivative can be written as well in the so-called Polyakov form---see, e.g., \cite{St_Kr_WL_cast} for a discussion of an alternative approach. Note that the derivation of the MM equations from the Schwinger-Dyson equations is grounded on the Mandelstam formula \begin{equation} \frac{\delta}{\delta \sigma_{\mu\nu} (x)} \ \Phi (\Gamma) = ig {\rm Tr} \[ F_{\mu\nu} \ \Phi (\Gamma_x) \] \end{equation} and/or on the Stokes theorem, so that the Wilson functionals which do not satisfy the corresponding restrictions (such as, e.g., cusped light-like loops) apparently cannot be straightforwardly treated within the same scheme. There are several other issues limiting the predictive power of the MM equations. Namely, there exists an interesting class of Wilson loops which possess very specific singularities originating, in particular, from the cusps and/or self-intersections of the contours and, in addition, from the light-like segments of the integration paths. The simplest example is given by a Wilson exponential evaluated along a cusped contour with two semi-infinite light-like sides, Fig. \ref{fig:0}. \begin{figure}[ht] $$\includegraphics[angle=90,scale=0.7]{cusp_1loop}$$ \vspace{0.0cm} \caption{\label{fig:0}The cusped integration contour on the light-cone with the one-gluon exchanges giving rise to the cusp anomalous dimension.} \end{figure} Already the leading order contribution to this Wilson exponential possesses all the peculiar singularities: the pure ultraviolet, the infrared (due to the infinite lengths of the sides), and the light-like cups divergences. This simple contour will arise in what follows as a building unit of many important Wilson loops and correlation functions. Physically it corresponds to the soft part of the factorized quark form factor, which has been studied in detail in \cite{WL_LC_rect, CAD_universal}. In the present work we propose and discuss a new approach to these issues, having in mind, as an instructive example, a very special type of Wilson loops---{planar rectangles} with light-like sides. Considerable interest to cusped light-like Wilson polygons has arisen thanks to the recently conjectured duality between the $n-$gluon planar scattering amplitudes in the ${\cal N} = 4 $ super-Yang-Mills theory and the vacuum average of {planar} Wilson loops formed, correspondingly, by $n$ light-like segments connecting space-time points ${x_i}$, so that their ``lengths'' $x_i - x_{i+1} = p_i$ are chosen equal to the external momenta of the $n-$gluon amplitude (see, e.g., \cite{WL_CFT} and references therein). It has been demonstrated that the infrared singularities of the former corresponds to the ultraviolet singularities of the latter, and the cusp anomalous dimension is the crucial constituent of the evolution equations \cite{KR87}. Wilson exponentials possessing light-like segments (or that are fully light-like) have been studied also in a different context \cite{WL_LC_rect}. The main observation is that the renormalization properties of these Wilson loops are more intricate than those of cusped Wilson loops defined on off-light-cone integration contours. Namely, the light-cone cusped Wilson loops are not multiplicatively renormalizable because of the additional light-cone singularities (besides the standard ultraviolet and infrared ones). It is possible, however, to construct a combined renormalization-group equation taking into account ultraviolet as well as light-cone divergences. The cusp anomalous dimension, which is the principal ingredient of this equation, is remarkably universal: it controls, e.g., the infrared asymptotic behavior of such important quantities as the QCD and QED Sudakov form factors, the gluon Regge trajectory, the integrated (collinear) parton distribution functions at large-$x$, the anomalous dimension of the heavy quark effective theory, etc. \cite{WL_LC_rect, KR87, CAD_universal, St_Kr_Pheno}. Another interesting field of application of cusped light-cone Wilson lines could be transverse-momentum dependent parton densities (TMDs) \cite{TMD_singular, CS_all}. The latter are {introduced} to describe the intrinsic transverse momentum of partons inside the nucleon, which is needed in the study of semi-inclusive processes within the (generalization of) the QCD factorization formalism \cite{TMD_singular, TMD_fact}. \section{Example: singularity structure of TMDs} Let us discuss the emergent singularities arising in TMDs beyond tree-approximation. At one-loop level, the following three classes of divergences appear: $(i)$ standard ultraviolet poles, which are removable by a normal renormalization procedure; $(ii)$ pure rapidity divergences, which depend on an additional rapidity cutoff, but do not violate renormalizability of TMDs; they can be resummed by means of the Collins-Soper evolution equation; $(iii)$ very specific overlapping divergences: they contain the ultraviolet and rapidity poles simultaneously and thus break down the standard renormalizability of TMDs. This situation resembles the problems with renormalizability of the light-like Wilson loops discussed above. However, the structure of Wilson lines is quite involved already in the tree-approximation. The most straightforward definition of ``a quark in a quark'' TMD, which meets the requirement of the {parton number interpretation}, reads \begin{eqnarray} & & {\cal F}_{\rm unsub.} \left(x, {\bm k}_\perp \right) = \frac{1}{2} \int \frac{d\xi^- d^2 {\xi}_\perp}{2\pi (2\pi)^2} \ {\rm e}^{-ik \cdot \xi} \cdot \nonumber \\ & & \times \left \langle p \ |\bar \psi_a (\xi^-, \bm{\xi}_\perp) {\cal W}_{n}^\dagger(\xi^-, \bm{\xi}_\perp; \infty^-, \bm{\xi}_\perp) {\cal W}_{\bm l}^\dagger(\infty^-, \bm{\xi}_\perp; \infty^-, {\infty}_\perp) \cdot \right. \nonumber \\ & & \left. \times \gamma^+ {\cal W}_{\bm l}(\infty^-, {\infty}_\perp; \infty^-, \bm{0}_\perp)_{\bm l} {\cal W}_{n}(\infty^-, \bm{0}_\perp; 0^-,\bm{0}_\perp)_{n} \psi_a (0^-,\bm{0}_\perp) | \ p \right \rangle \ \label{eq:general} \end{eqnarray} with ${\xi^+=0}$. Here we define the semi-infinite Wilson lines evaluated along a four-vector $w$ as $$ {\cal W}_w(\infty; \xi) \equiv {\cal P} \exp \left[ - i g \int_0^\infty d\tau \ w_{\mu} \ A_{a}^{\mu}t^{a} (\xi + w \tau) \right] \ , $$ where the vector $w$ can be light-like $w_L = n^\pm\ , \ (n^\pm)^2 =0$, or transverse $w_T = {\bm l}$. Formally, the integration of (\ref{eq:general}) over $\bm k_\perp$ is expected to give the collinear (also called integrated) PDF \begin{equation} \int\! d^2 \bm k_\perp \ {\cal F}_{\rm unsub.} (x, \bm k_\perp) = \frac{1}{2} \int \frac{d\xi^- }{2\pi } \ {\rm e}^{-ik^{+}\xi^{-} } \ \left\langle p\ |\bar \psi_a (\xi^-, \bm 0_\perp){\cal W}_n(\xi^-, 0^-) \gamma^+ \psi_a (0^-,\bm 0_\perp) | \ p \right\rangle \ = f_a(x) \ . \label{eq:u_to_i} \end{equation} However, this is only justified in tree approximation. It is worth noting that the normalization of the above TMD \begin{equation} {\cal F}_{\rm unsub.}^{(0)} (x, {\bm k}_\perp) = \frac{1}{2} \int \frac{d\xi^- d^2 \bm{\xi}_\perp}{2\pi (2\pi)^2} {\rm e}^{- i k^+ \xi^- + i \bm{k}_\perp \cdot \bm{\xi}_\perp} { \langle p \ | }\bar \psi (\xi^-, \bm{\xi}_\perp) \gamma^+ \psi (0^-, \bm 0_\perp) { | \ p \rangle } = \delta(1 - x ) \delta^{(2)} (\bm k_\perp) \ \label{eq:tree_tmd} \end{equation} can be most easily obtained by making use of the {canonical quantization procedure in the light-cone gauge}, where longitudinal Wilson lines become equal to unity and where equal-time commutation relations for creation and annihilation operators $\{a^\dag (k, \lambda), a(k, \lambda)\}$ immediately yield the parton number interpretation \begin{equation} {\cal F}_{\rm unsub.}^{(0)} (x, {\bm k}_\perp) \sim \langle\ p \ | \ a^\dag(k^+, \bm k_\perp; \lambda) a(k^+, \bm k_\perp; \lambda) \ | \ p\ \rangle \ . \label{eq:parton_N} \end{equation} The usage of ``tilted'' gauge links in the operator definition of TMDs does not meet this requirement. We visualize the geometrical layout of various Wilson lines in the operator definition of TMDs in Fig. \ref{fig:geo1}, \ref{fig:geo2}, \ref{fig:geo3} and discuss relevant issues in their captions. \begin{figure}[p] \includegraphics[width=0.45\textwidth,height=0.70\textheight,angle=90]{wilson_lines_PDF.eps}\hspace{2pc}% \caption{\label{fig:geo1} Geometry of the contours in unsubtracted TMDs with light-like (upper panel) and off-light-cone (lower panel) longitudinal Wilson lines and their reduction to integrated PDFs in tree approximation. In the former case, the transverse Wilson lines vanish after $\bm k_\perp$-integration, while the longitudinal Wilson lines turn into an one-dimensional connector ${\cal W}_n(\xi^-, 0^-)$. In the off-light-cone schemes, the mutual compensation of transverse Wilson lines at infinity is not visible. Moreover, the integrated configuration contains two non-vanishing off-light-cone Wilson lines, which apparently are not equivalent to the collinear connector ${\cal W}_n(\xi^-, 0^-)$. The interrogation marks next to the transverse Wilson lines symbolize the lacking of any consistent treatment in TMD formulations with off-light-cone (shifted) Wilson lines. In contrast, the transverse Wilson lines appear naturally in ``light-cone'' schemes.} \end{figure} \begin{figure}[p] \includegraphics[width=0.45\textwidth,height=0.70\textheight,angle=90]{wilson_lines_SF.eps} \caption{\label{fig:geo2}Comparative layout of Wilson lines in unsubtracted soft factors and visualization of the reduction to the collinear case. The upper panel shows the soft factor in momentum space, as proposed in Refs. \cite{CS_all}. The lower panel presents the tilted off-light-cone integration paths in impact parameter space, as well as the result of the reduction to the collinear $\bm b_\perp \to 0$ configuration.} \end{figure} \begin{figure}[p] \includegraphics[width=0.45\textwidth,height=0.70\textheight,angle=90]{wilson_lines_full_SF.eps} \caption{\label{fig:geo3}Comparative layout of Wilson lines in subtracted soft factors. The upper panel corresponds to the soft factor of the TMD distribution function which enters the factorization with pure light-like Wilson lines. The lower panel {has the same setup, but with} the longitudinal Wilson lines shifted off the light-cone.} \end{figure} Beyond tree-approximation, the virtual diagrams producing terms with overlapping singularities are shown in Fig.\ \ref{fig:1}. The typical extra divergency stems from the one-loop vertex-type graph Fig. \ref{fig:1}(a) in covariant gauges or from the self-energy graph Fig. \ref{fig:1}(b) in the light-cone gauge (in the large-$N_c$ limit) and reads \begin{equation} {\rm TMD}_{\rm UV\otimes LC} = - \frac{\alpha_s N_c}{2\pi} \Gamma (\epsilon) \left[ 4\pi \frac{\mu^2}{-p^2} \right]^{\epsilon} \ \delta (1-x) \delta^{(2)} (\bm{k}_\perp ) \ {\int_0^1\! dx \ \frac{x^{1-\epsilon}}{(1-x)^{1+\epsilon}} } \ . \end{equation} The standard ultraviolet pole in the Gamma-function $\Gamma (\epsilon)$ is accompanied by an additional singularity in the integral. The latter is due to the integration over infinite gluon rapidity and cannot be treated by dimensional regularization, calling for an extra (rapidity) cutoff. The reason for renormalizability violation in the leading order contribution to TMDs is that light-like Wilson lines (or the ``standard'' quark self-energy in light-cone gauge) produce more singular terms than usual Green functions do. \begin{figure}[ht] $$\includegraphics[angle=90,scale=0.7]{wl_lc_1loop}$$ \vspace{0.0cm} \caption{\label{fig:1}The virtual one-loop Feynman graphs which produce extra singularities: $(a)$---vertex-type fermion-Wilson line interaction in covariant gauge; $(b)$---self-energy graph which yields the extra divergency in light-cone gauge; $(c,d)$ are the counter-parts of $(a,b)$ from the soft factor made of Wilson lines.} \end{figure} To solve the problems with extra singularities and renormalizability in TMDs, a variety of (possibly non-equivalent) methods has been proposed. Working in the covariant Feynman gauge, Ji, Ma and Yuan proposed a scheme which utilizes tilted (off-light-cone) longitudinal Wilson likes directed along the vector $n_B^2 \neq 0$ \cite{JMY}. Transverse Wilson lines at the light-cone infinity cancel in covariant gauges, while the rapidity cutoff $\zeta = (2 p\cdot n_B)^2/|n_B^2|$ marks the deviation of longitudinal Wilson lines from a pure light-like direction. A subtracted soft factor then contains non-light-like Wilson lines as well. Obviously, such off-light-cone unsubtracted TMDs with the light-like vector $n^-$ replaced by the vector $n_{\rm B} = (-{\rm e}^{2y_B}, 1, \bm 0_\perp)$ do not obey the equation (\ref{eq:u_to_i}), not even {at tree level}. However, it is possible to formulate a ``secondary factorization'' method which allows one to express off-light-cone TMDs (in impact parameter space ${\cal F} (x, \bm b_\perp)$) as a convolution of integrated PDFs and perturbative coefficient functions in the perturbative region (that is, at small $\bm b_\perp$), see \cite{JMY}. In publications \cite{CS_all} it was proposed to explore the renormalization-group properties of unsubtracted TMDs (\ref{eq:general}) and to make use of their anomalous dimension as a tool to discover the {minimal} layout of Wilson lines in the soft factor that provides a cancelation of overlapping dependent terms. It has been demonstrated (in the leading $O(\alpha_s)$-order) that the extra contribution to the anomalous dimension is exactly the cusp anomalous dimension \cite{KR87}, which is a crucial element of the investigation of non-renormalizible cusped light-like Wilson loops. Making use of specially chosen soft factors, one can get rid of the extra divergences in the operator definition of the TMDs, however paying a price in the form of significant complication of the structure of the Wilson lines in the above definition. In the present work we discuss another approach to the problems of light-cone cusped Wilson loops \cite{ChMVdV_2012}. To this end, it appears instructive to study those properties shared by such apparently different quantities as TMDs, light-like Wilson polygons, etc., which originate in their light-cone structure and arise in the form of the ``too singular'' non-renormalizable terms. \section{Schwinger dynamical principle and area evolution for smooth Wilson loops} We made use of the observation that in the large-$N_c$ limit, in the transverse null-plane, for the light-like {planar} {dimensionally regularized (not renormalized)} Wilson rectangles, the area derivatives introduced in the previous sections can be reduced to the normal ones. The area variational equations in the coordinate representation describe the evolution of light-like Wilson polygons and represent, therefore, the ``equations of motion'' in loop space, valid for a specific class of its elements. As a result, the obtained differential equations give us a closed set of dynamical equations for the loop functionals, and can in principle be solved in several interesting cases. Let us start with the {quantum dynamical principle} proposed by Schwinger \cite{Schwinger51}: the action operator $S$ defines variations of arbitrary quantum states, so that \begin{equation} { \delta \langle \ \alpha' \ | \ \alpha''\ \rangle } = \frac{i}{\hbar} { \langle \ \alpha' \ | \delta S | \ \alpha'' \ \rangle } \ . \label{eq:principle} \end{equation} The area variations (\ref{eq:area_derivative}) of field exponentials $\Phi(\Gamma)$ yield \begin{equation} { \frac{\delta}{\delta \sigma} \langle \ \alpha' \ | \Phi(\Gamma) | \ \alpha''\ \rangle } = \frac{i}{\hbar} { \langle \ \alpha' \ | \frac{\delta \hat S}{\delta \sigma} \Phi(\Gamma) | \ \alpha'' \ \rangle } \ , \label{eq:principle_mod} \end{equation} where $\hat S$ is yet to be defined. The loop space consists of scalar objects with different topological and geometrical features, hence the equations of motion in this space must be the laws which state how those objects change their shape. It means that ``motion'' in loop space is equivalent to variation of the shapes of integration contours in the Wilson loops \cite{MM_WL}. Therefore, we have to find the proper operator $\hat S$, which governs the shape variations of the light-like cusped loops (Wilson planar polygons). Following the standard method, one makes use of Eq. (\ref{eq:principle}) in the form (\ref{eq:sch_dy_YM}) and gets the system of the MM Eqs. (\ref{eq:MM_general}). Alternatively, we will try to get rid of the operations which tacitly assume the property of smoothness of the Wilson loops of interest. Consider, e.g., a generic Wilson loop $W (\Gamma)$ without mentioning if it is smooth or not. The two leading terms of its perturbative series are given by \begin{equation} {\cal W} (\Gamma) = {\cal W}^{(0)} + {\cal W}^{(1)} = 1 - \frac{g^2 C_F}{2} \ \oint_\Gamma \oint_\Gamma \ dz_\mu dz_\nu' \ D^{\mu\nu} (z - z') + O(g^4) \ , \nonumber \label{eq:W_generic_pert} \end{equation} where $D^{\mu\nu}$ is the dimensionally regularized ($\omega = 4 - 2 \epsilon$) free gluon propagator \begin{equation} D^{\mu\nu} = - g^{\mu\nu} \ \Delta (z - z') \ , \ \Delta(z-z') = \frac{\Gamma(1-\epsilon)}{4\pi^2} \ \frac{(\pi \mu^2)^\epsilon}{[- (z-z')^2 + i0]^{1- \epsilon}} \ . \label{eq:gluon_prop_DR} \end{equation} For convenience's sake, we work in the Feynman covariant gauge and separate out the scalar part of the propagator $\Delta (z)$. The issues related to gauge- and regularization independence of the calculations will be considered elsewhere. Then, if the l.h.s. of the Eq. (\ref{eq:principle_mod}) acts on the Wilson exponential (\ref{eq:W_generic_pert}), one gets \begin{equation} \frac{\delta {\cal W} (\Gamma)}{\delta \sigma_{\mu\nu}} = \frac{g^2 C_F}{2} \ \frac{\delta }{\delta \sigma_{\mu\nu}} \oint_\Gamma \oint_\Gamma \ dz_\lambda dz^{'\lambda} \ \Delta (z - z') + O(g^4) \ . \label{eq:W_area_diff_2} \end{equation} Using the Stokes theorem (let us assume for a moment that we are allowed to do so), we obtain \begin{equation} \oint_\Gamma dz_\lambda \ {\cal O}^\lambda = \frac{1}{2} \int_\Sigma \ d\sigma_{\lambda \rho} (\partial^\lambda {\cal O}^\rho - \partial^\rho {\cal O}^\lambda) \ , \ {\cal O}^\lambda = \oint_\Gamma dz_\lambda' \ \Delta (z) \ , \label{eq:W_diff_Stokes} \end{equation} where $\Gamma$ is considered as the boundary of the surface $\Sigma$. We get then the leading perturbative term of the MM equation (\ref{eq:MM_general}): \begin{equation} \partial_\mu \frac{\delta {\cal W}(\Gamma_\bigcirc) }{\delta \sigma_{\mu\nu} (x)} = \frac{g^2N_c}{2} \oint_{\Gamma_\bigcirc} \ dy_\nu \ \delta^{(\omega)} (x -y) + O(g^4)\ . \label{eq:MM_LO_smooth} \end{equation} We must treat this result with due caution: in the course of the derivation, we assumed that the Stokes theorem is valid for all Wilson loops which we consider. However, the last statement is not true in general, that is why we mark the ``good'' contours with a special index $\Gamma_\bigcirc$. It is interesting that in $2D$ QCD, the area differentiation turns into the ordinary derivative, by virtue that the gluon propagator (\ref{eq:gluon_prop_DR}) for $\omega=2$ gets logarithmic in $z$: \begin{equation} {\cal W}(\Gamma_\bigcirc)^{\rm 2D} = \exp\[ - \frac{g^2 N_c}{2} \Sigma \] \ , \ \Sigma = \ {\rm area \ inside\ } \Gamma_\bigcirc \ , \label{eq:2D_area} \end{equation} so that $ 2 \ln W'_\Sigma = - {g^2 N_c}. $ Evaluating, in the same way, the NLO terms, one obtains the full MM Eq. (\ref{eq:MM_general}). However, we interrupt here and return a bit backward, since we are mostly interested in those Wilson functionals that do not satisfy (or, at least, do not satisfy straightforwardly) the conditions of the applicability of the Stokes theorem. We will, therefore, continue with the study of the shape variations of Wilson loops without relying upon the Stokes theorem, but keeping in mind an explicit form of the free gluon propagator (which possesses a specific light-cone/rapidity divergence), see Eq. (\ref{eq:gluon_prop_DR}). \section{Singularities of Wilson rectangles} We are now in a position to extend the Schwinger approach to a more complicated case and to try to derive the corresponding area evolution equations. The calculation of cusped light-cone Wilson loops beyond tree approximation in different gauges and the justification of gauge independence calls for a careful treatment of a variety of divergences already in leading order. Special attention must by paid to the separation of the rapidity divergences and the standard ultraviolet ones \cite{WL_RG, WL_LC_rect, WL_LC_rapidity}. In the 't Hooft (large-$N_c$) limit one obtains \cite{WL_LC_rect} \begin{eqnarray} & & W(\Gamma_\Box) = 1 - \frac{1}{\epsilon^2}\ \frac{\alpha_s N_c}{2\pi} \ \(\[{-2 N^+ N^-\mu^2 + i0}\]^\epsilon + \[{2 N^+ N^-\mu^2 + i0}\]^\epsilon \) \\ & & + \frac{\alpha_s N_c}{2\pi} \( \frac{1}{2} \ln^2 \frac{N^+N^-+ i0}{-N^+N^-+i0} + {\rm finite\ terms} \) + O(\alpha_s^2) \ , \nonumber \label{eq:WL_LC_1loop} \end{eqnarray} with the Mandelstam variables in the momentum space, ${ s = (p_1 + p_2)^2}$ and ${t = (p_2 + p_3)^2}$, map onto the {area} variables in the coordinate space, so that ${ s/2} = {- t/2} \to N^+ N^-$. We will show separately that the result (\ref{eq:WL_LC_1loop}) is not only gauge invariant, but is independent of any regularization of light-cone and ultraviolet divergences and of the way they are separated. The latter issue is of considerable importance to the study of the operator structure of transverse-momentum dependent parton densities, the jet and soft functions in the soft-collinear effective theory, the infrared properties of the high-energy scattering amplitudes, etc. (see, e.g., \cite{CS_all, SCET_TMD} and references therein). The transverse null-plane is defined by the condition $\vecc z_\perp = 0$; therefore, the shape variations (which give rise to the infinitesimal area changes) are defined as \begin{equation} { \delta \sigma^{+-} } = { N^+ \delta N^- } \ , \ { \delta \sigma^{-+} } = - { N^- \delta N^+ } \ . \label{eq:delta_area} \end{equation} We assume that these operations do make sense only at the corner points ${x_i}$, and introduce the ``left'' and ``right'' variations, as shown in Fig. \ref{fig:2}. \begin{figure}[ht] $$\includegraphics[angle=90,width=0.7\textwidth]{wl_lc_area}$$ \caption{Infinitesimal area transformations for a light-cone rectangle on the null-plane: we consider only those area variations that conserve the angles between the sides. These variations are defined at the corners $x_i$.} \label{fig:2} \end{figure} Rectangular planar Wilson loop ${ W(\Gamma_{\Box})}$ is known to lack multiplicative renormalizability \cite{WL_LC_rect}. In order to decrease the power of singularity that violates the renormalizability, one can follow the scheme proposed in \cite{CAD_universal}. Having in mind Eq. (\ref{eq:delta_area}), we define the area logarithmic derivative \begin{equation} \frac{\delta}{\delta \ln \sigma} \equiv \sigma_{+-} \frac{\delta}{\delta \sigma_{+-}} + \sigma_{-+} \frac{\delta}{\delta \sigma_{-+}} \ \label{eq:area_log} \end{equation} and act with this operator on the r.h.s. of the Eq. (\ref{eq:WL_LC_1loop}): \begin{equation} \frac{\delta}{\delta \ln \sigma} \ \ln W(\Gamma_\Box) = - { \frac{\alpha_s N_c}{2\pi} } \ \frac{1}{\epsilon} \ \( \[{-2N^+N^-\mu^2 + i0}\]^\epsilon + \[{2N^+N^-\mu^2 + i0}\]^\epsilon \) \ . \label{eq:log_der} \end{equation} Then, the result is finite (after additional logarithmic differentiation in the ultraviolet scale $\mu$) and is given by the {cusp anomalous dimension} \begin{equation} \mu \frac{d}{d\mu} \frac{\delta \ \ln W(\Gamma_\Box)}{\delta \ln \sigma} = - 4 \ { \Gamma_{\rm cusp} } \ , \ {\Gamma_{\rm cusp} = \frac{\alpha_s N_c}{2 \pi} } + O(\alpha_s^2) \ . \label{eq:full_der} \end{equation} Therefore, we have obtained the result (\ref{eq:full_der}) given that the infinitesimal shape variations are defined as in (\ref{eq:area_derivative}). Eq.(\ref{eq:full_der}) describes then the dynamics of the cusped planar light-like Wilson loops \cite{ChMVdV_2012}. We established, therefore, the connection between the {geometry} of the loop space (in terms of the area/shape differentials) and the {dynamical properties} of the fundamental degrees of freedom---the gauge- and regularization-invariant planar {light-like Wilson loops}. \section{Combined evolution from the Schwinger principle} The very possibility to obtain a finite result by means of Eqs. (\ref{eq:log_der}, \ref{eq:full_der}) is a direct consequence of the geometrical properties of loop space, when considering the non-renormalizable cusped light-like Wilson loops. To show this explicitly, we restrict ourselves to shape variations (\ref{eq:delta_area}), and apply the area derivative to the planar Wilson rectangle \begin{eqnarray} & & \frac{\delta W(\Gamma_\Box) }{\delta \sigma_{\mu\nu}} = \frac{g^2C_F}{2} \frac{\Gamma(1-\epsilon) (\pi \mu^2)^\epsilon}{4\pi^2} \ \frac{\delta }{\delta \sigma_{\mu\nu}} \sum_{i,j} (v_j^\lambda v_j^\lambda) \cdot \nonumber \\ & & \times \int_0^1\int_0^1\! \frac{d\tau d\tau'}{[- (x_i - x_j - \tau_i v_i + \tau_j v_j)^2 + i0]^{1 -\epsilon}} \ , \label{eq:WL_rect_area} \end{eqnarray} where the sides are parameterized as $z_i^\mu = x_i^\mu - v_i^\mu \tau$ with the vectors $v_i$ having the dimension $[{\rm mass}^{-1}]$ \cite{WL_LC_rect}. A peculiar feature of the planar light-like contours is that the area dependence separates out from the integrals and can be calculated explicitly (taking into account that $2 (v_1v_2) = 2 N^+N^-$, see Eq. (\ref{eq:delta_area})) \begin{equation} W^{(1)}(\Gamma_\Box) = - \frac{\alpha_s N_c}{2\pi} {\Gamma(1-\epsilon)} (\pi \mu^2)^\epsilon \ (- 2 N^+N^-)^\epsilon\ \frac{1}{2} \int_0^1\int_0^1\! \frac{d\tau d\tau'}{[(1-\tau)\tau']^{1 -\epsilon}} \ . \label{eq:WL_rect_area_NN} \end{equation} Moreover, light-like Wilson lines with $v_i^2 = 0$ develop an extra singularity, which shows up in the form of a second-order pole $\sim \epsilon^{-2}$, while the cusps violate conformal invariance of the Wilson loop because the ``skewed'' scalar products $(v_i v_j) \neq 0$ replace the conformal ones $v_i^2$. Hence, performing the area $\delta / \delta \ln \sigma = \delta / \delta \ln (2 N^+N^-) $ and the UV-scale logarithmic differentiation of Eq. (\ref{eq:WL_rect_area_NN}) and summing up all relevant terms, we obtain \begin{equation} \mu \frac{d }{d \mu } \ {\[ \frac{\delta}{\delta \ln \sigma} \ \ln \ W (\Gamma) \] } = - \sum { \Gamma_{\rm cusp} } \ , \label{eq:mod_schwinger} \end{equation} which has been foreseen in Eq. (\ref{eq:full_der}) and which is derived now directly from the Schwinger quantum dynamical principle. It is natural that this result is akin, in some sense, to the situation in $2D$ QCD mentioned above. The area derivative becomes the ordinary derivative for the same reason: one has effectively planar Wilson loops, thus the MM Eqs. becomes a closed system and, in principle, solvable \cite{MM_WL, WL_Renorm}. Let us emphasize that the r.h.s. of Eq. (\ref{eq:mod_schwinger}) is regulated by the cusp anomalous dimension, which is a universal quantity that is independent of the shape of particular contour and which is known perturbatively up to the $O(\alpha_s^3)$ order. It is therefore worth analyzing if the above result is only a leading order approximation, or if it also will be valid for higher order approximations. Let us take into consideration the property of linearity of the (angular-dependent) cusp anomalous dimension in the infinitely large angle asymptotics with respect to the logarithm of the cusp angle $\chi \to \frac{1}{2}\ln \frac{(2 v_i v_j)^2}{v_i^2 v_j^2}$ \cite{KR87}: \begin{equation} \lim_{\chi \to \infty} \Gamma_{\rm cusp} (\chi, \alpha_s) = \sum \alpha_s^n C_n (W) a_n (W) \ \ln \frac{(2 v_i v_j)}{|v_i| |v_j^2|} \ , \label{eq:cuspCD_LC} \end{equation} with the ``maximally non-Abelian'' coefficients given by \begin{equation} {C_k} \sim C_F \ N_c^{k-1} \to {\frac{N_c^k}{2} } \ , \end{equation} and $a_n$ are cusp-independent factors. This asymptotic regime coincides with the light-cone case with the angular-dependent logarithms being transformed into extra pole terms in $\epsilon$: $\chi \to \frac{(v_iv_j)^\epsilon}{\epsilon}$, see \cite{WL_LC_rect, KR87}. More specifically, the area variable $ \ \sim (v_iv_j)$ enters the regularized cusp anomalous dimension in the light-like limit as \begin{equation} \Gamma_{\rm cusp} ({\rm area}, \epsilon, \alpha_s) = \sum \alpha_s^n C_n (W) a_n (W) \ \frac{{\rm area}^\epsilon}{\epsilon} \ , \label{eq:cuspCD_LC_1} \end{equation} and, after logarithmic differentiation, one obtains a perturbative expansion of the cusp anomalous dimension \begin{equation} \lim_{\epsilon \to 0} \frac{d \ \Gamma_{\rm cusp} ({\rm area}, \epsilon, \alpha_s)}{d\ \ln {\rm area}} = \sum \alpha_s^n C_n (W) a_n (W) \ , \label{eq:cuspCD_LC_2} \end{equation} which justifies the previous result (\ref{eq:mod_schwinger}) in the NLO by virtue that $$ \Gamma_{\rm cusp} = - d \ln W / d \ln \mu \ . $$ This implies that the result (\ref{eq:mod_schwinger}) is, in fact, an all-order one, akin to the MM Eq. (\ref{eq:MM_general}): they both are exact and non-perturbative, while the r.h.s's of each one can be calculated perturbatively. It is worth noting that Eq. (\ref{eq:full_der}) agrees with the non-Abelian exponentiation theorem for the dimensionally regularized Wilson loops \begin{equation} W (\Gamma_\Box; \epsilon) = \exp \[ \sum_{k=1} \alpha_s^k \ C_k (W) F_k (W) \ \] \ , \end{equation} with the summation going over all two-particle irreducible diagrams, whose contribution is given by the so-called ``web'' functions $F_k$ \cite{WL_expo, KR87}. Thus, Eq. (\ref{eq:full_der}) can be used in calculation of the higher-order terms in the cusp anomalous dimension, by virtue that one has a closed recursive system of the perturbative equations. Now we can apply the methodology described above to the TMDs with the light-like longitudinal Wilson lines ${\cal F} (x, \vecc k_\perp)$, Eq. (\ref{eq:general}), what yields \begin{equation} \mu \frac{d }{d \mu } \ \[ \frac{d}{d \ln \theta} \ \ln \ { {\cal F} (x, \vecc k_\perp) } \] = 2 { \Gamma_{\rm cusp} } \ , \label{eq:tmd_combined} \end{equation} with the ``area'' being encoded in the rapidity cutoff parameter $\theta \sim (p N^-)^{-1}$ \cite{CS_all}. Another interesting example is provided by the $\Pi$-shape semi-loop with one finite light-like segment \cite{WL_Pi}. In the one-loop order, we have in the large-$N_c$ limit \begin{eqnarray} & & W(\Gamma_\Pi) = 1 + \frac{\alpha_s N_c}{2\pi} \ + \[ - L^2 (NN^-) + L (NN^-) - \frac{5 \pi^2}{24} \] \ , \\ & & L(NN^-) = \frac{1}{2}\(\ln (\mu N N^- + i0) + \ln (\mu NN^- + i0) \)^2 \ , \nonumber \label{eq:pi_1loop} \end{eqnarray} where the area is given by the product of the light-like $N^-$ and non-light-like $N$ vectors, see Fig. \ref{fig:5}. \begin{figure}[ht] $$\includegraphics[angle=90,scale=0.7]{wl_lc_pi_shape}$$ \vspace{0.2cm} \caption{\label{fig:5}$\Pi$-shape integration contour and the infinitesimal area variations.} \end{figure} The $\Pi$-shaped Wilson loop (\ref{eq:pi_1loop}) also satisfies Eq. (\ref{eq:mod_schwinger}): \begin{equation} \mu \frac{d }{d \mu } \ \[ \frac{d}{d \ln \sigma} \ \ln \ { W(\Gamma_\Pi) } \] = - 2 { \Gamma_{\rm cusp} } \ , \end{equation} the latter controls for the renormalization-group evolution of the integrated PDFs in the large-$x$ limit, as well as the anomalous dimensions of conformal operators with large Lorentz spin \cite{WL_Pi}. The $\Pi$-shape contour can be split and moved apart to separate two planes by the transverse distance $\bm \xi_\perp$. The Wilson loop obtained in such a way is expected to be ``dual'' to the TMD. This duality implies that both the double-planar $\Pi$-shaped Wilson loop, Fig. \ref{fig:7}, and the TMD (\ref{eq:general}) obey the same combined evolution equation (\ref{eq:tmd_combined}). The detailed analysis of this configuration will be presented elsewhere. \begin{figure}[ht] $$\includegraphics[angle=90,width=0.7\textwidth]{tmd_dual}$$ \caption{Conjectured ``dual'' Wilson loop having the combined evolution similar to the one of a TMD. Transverse Wilson lines are not shown for simplicity.} \label{fig:7} \end{figure} \section{Conclusions and outlook} The universal quantum dynamical principle by Schwinger provides a proper approach to the description of the dynamics of the loop space. The gauge invariant Wilson loops are considered then as the only degrees of freedom, and the Makeenko-Migdal equations (\ref{eq:MM_general}) stem from the Schwinger-Dyson equations applied to the renormalizable loops. In general case, the system of the MM Eqs. is not closed and cannot be immediately used in practical calculations in QCD. \begin{figure}[ht] $$\includegraphics[angle=90,scale=0.7]{wl_lc_general}$$ \caption{\label{fig:3}Generic infinitesimal area variations responsible for the conjectured quantum-dynamical loop equations for light-like Wilson $n-$polygons. Evaluation of minimal surface differentials for more complicated cusped Wilson loops is required to derive corresponding area evolution equations based on the quantum dynamical principle \cite{WL_min_surf}.} \end{figure} In the present work we showed that it is possible to design a relevant system of equations of motion valid for the cusped planar light-like Wilson loops, taking into account that the latter possess a very specific singularity structure compared to their off-light-cone analogues. General solution of this problem has not been found yet, but we have managed to demonstrate that some simplifications make it possible to propose a new potentially fruitful method to deal with such Wilson exponentials. Namely, in the large-$N_c$ limit, the {planar} rectangular light-like contours at $\vecc z_\perp =0$ enable us to reduce the area functional derivative to the normal derivative for dimensionally regularized Wilson loops. As the result, the equations which describe the infinitesimal shape variations in coordinate space appear to be dual to the energy evolution equations for cusped Wilson loops in the space of the Mandelstam momentum variables. Within the framework we proposed, the dynamics of elements of loop space is introduced by means of obstructions of the initially smooth Wilson loops, which play, therefore, the role of {\it sources} in Schwinger's ``fields and sources'' picture. We have argued, therefore, that the Schwinger quantum dynamical principle can be used as an effective tool to study (at least) one important class of the elements of loop space---the cusped planar Wilson polygons on the light-cone. We implemented the program only in one of the simplest situations, the planar rectangle. In Fig. \ref{fig:3}, a more involving configuration is visualized, the arbitrary quadrilateral path, the area evolution of which is far from being trivial and deserves a separate study. \section{Acknowledgements} We appreciate stimulating discussions and useful critical remarks by I.V. Anikin, Y.M. Makeenko and N.G. Stefanis. We thank the participants of the JLab QCD Evolution workshop, the summer meetings in ECT*, Trento, and the joint seminar at the University of Antwerp for useful discussions.
\section{Order routing and order placement in electronic financial markets} The trading process in today's automated financial markets is divided into several stages, each taking place on a different time horizon: portfolio allocation decisions are usually made on a monthly or daily basis and translate into trades that are executed over time intervals of several minutes to several days through streams of {\it orders} placed at high frequency, sometimes hundreds in a single minute \citep{Cont2011}. Existing studies on optimal trade execution \cite{bertsimas98,almgren00} have investigated how the execution cost of a large trade may be reduced by splitting it into multiple {\it orders} spread in time. Once this {\it order scheduling } decision is taken, one still needs to specify how each individual order should be {\it placed}: this order placement decision involves the choice of an {\it order type} (limit order or market order), order size and destination, when multiple trading venues are available. For example, in the U.S. equity market there are more than ten active exchanges where a trader can buy or sell the same securities. We focus here on this {\it order placement} problem: given an order which has been scheduled, deciding what type of order --market or limit order-- and which trading venue to submit it to. Orders are filled over short time intervals of a few milliseconds to several minutes and the mechanism through which orders are filled in the limit order book are relevant for such order placement decisions. When trading large portfolios, market participants need to make such decisions repeatedly, thousands of times a day, and their outcomes have a large impact on each participant's transaction cost as well as on aggregate market dynamics. Order placement and order routing decisions play an important role in financial markets. Brokers are required by law to deliver the best execution quality to their clients and empirical evidence confirms that a large percentage of market orders in the U.S. and Europe is sent to trading venues providing lower execution costs or smaller delays \cite{Boehmer2006,Foucault2008}. Market orders gravitate towards exchanges with larger posted quote sizes and low fees, while limit orders are submitted to exchanges with high rebates and lower execution waiting times (see \cite{Moallemi2011}). These studies demonstrate how investors' aggregate order routing decisions have a significant influence on market dynamics, but a systematic study of the order routing problem from the investor's perspective is lacking. A reduced-form model for routing an infinitesimal limit order to a single destination is used by \cite{Moallemi2011}, while \cite{Ganchev2010} and \cite{Laruelle2009} propose numerical algorithms to optimize order executions across multiple dark pools, where supply/demand is unobserved. To the best of our knowledge this work is the first to provide a detailed treatment of investor's order placement decision in a multi-exchange market. Early work on optimal trade execution \cite{bertsimas98,almgren00} did not explicitly model the process whereby each order is filled, but more recent formulations have tried to incorporate some elements in this direction. In one stream of literature (see \cite{obizhaeva05}, \cite{Alfonsi2010}, \cite{Predoiu2011}) traders are restricted to using market orders whose execution costs are given by an idealized order book shape function. Another approach has been to model the process through which an order is filled as a dynamic random process \citep{Cont2011,Cont2011b} leading to a formulation of the optimal execution problem as a stochastic control problem: this formulation has been studied in various settings with limit orders (\cite{Bayraktar2011}, \cite{Gueant2012}) or limit and market orders \cite{Guilbaud2012,Huitema2012} but its complexity makes it computationally intractable unless restrictive assumptions are made on price and order book dynamics. In the present work, we adopt a simpler, more tractable approach by separating the {\it order placement} decision from the scheduling decision: assuming that the trade execution schedule has been specified, we focus on the task of filling the scheduled batch of orders by optimally distributing it across trading venues and order types. Decoupling the scheduling problem from the order placement problem leads to a more tractable approach which is closer to market practice and allows us to incorporate some realistic features which matter for order placement decisions, while conserving analytical tractability. Our key contribution is a quantitative formulation of the order placement problem which takes into account multiple important factors - the size of an order to be executed, lengths of order queues across exchanges, statistical properties of order flows in these exchanges, trader's execution preferences, and the structure of liquidity rebates across trading venues. Our problem formulation is tractable, intuitive and blends the aforementioned factors into an optimal allocation of limit orders and market orders across available trading venues. Order routing heuristics employed in practice commonly depend on past order fill rates at each exchange and are inherently backward-looking. In contrast, our approach is forward-looking - the optimal order allocation depends on current queue sizes and distributions of future trading volumes across exchanges. When only a single exchange is available for execution, this order placement problem reduces to the problem of choosing an optimal split between market orders and limit orders. We derive an explicit solution for this problem and analyze its sensitivity to the order size, the trader's urgency for filling the order and other factors. In a case of multiple exchanges we also derive a characterization of the optimal order allocation across trading venues. Finally, we propose a fast and flexible numerical method for solving the order placement problem in a general case and demonstrate its efficiency through examples. Our numerical examples show that the use of our optimal order placement method allows to substantially decrease trading costs in comparison with some simple order placement strategies. An important aspect of our framework is to account for the {\it execution risk}, the risk of not filling an order, by penalizing such outcomes. This is different from most other studies which focus on the risk of price variations over the course of a trade execution \cite{almgren00,huberman2005} but assume that orders are always filled, or studies that ignore execution risk altogether \cite{bertsimas98,Bayraktar2011}. In our framework the penalty for execution risk plays an important role. When it is costly to catch up on the unfilled portion of the order, the optimal allocation shifts from limit to market orders. Although market orders are executed at a less favorable price, their execution is more certain and it becomes optimal to use them when the execution risk is a primary concern. Optimal limit order sizes are strongly influenced by total quantities of orders queueing for execution at each exchange and by distributions of order outflows from these queues. For example, if the queue size at one of the exchanges is much smaller than the expected future order outflow there, it is optimal to place a larger limit order on that exchange. \cite{Moallemi2011} argue that such favorable limit order placement opportunities vanish due to competition and strategic order routing of individual traders. However the empirical results in that study also show that short-term deviations from the equilibrium are a norm, and can therefore be exploited in our optimization framework to improve limit order placement decisions. Finally, we find that the targeted execution size plays an important role - limit orders are used predominantly to execute small trades and market orders are used to execute larger quantities, as long as their cost is less than the penalty for falling behind the target. This is a due to the fact that the amount of limit orders that can be realistically filled at each exchange is naturally constrained by the corresponding queue size and the order outflow distribution, so to execute larger quantities the trader needs to rely on market orders. We find that the optimal order allocation almost always sends limit orders to all available exchanges in an attempt to diversify execution risk, which suggests a benefit in having multiple exchanges. However, when order flows on different exchanges are highly positively correlated, these diversification advantages fade and the cost of execution becomes higher than on a single consolidated exchange. Section \ref{sec.routing} describes our formulation of the order placement problem and presents simple and intuitive conditions for the existence of an optimal order placement. In Section \ref{sec.1exchg} we derive an optimal split between market and limit orders for a single exchange. Section \ref{sec.2exchg} analyzes the general case of order placement on multiple trading venues. Section \ref{sec.sa} presents a numerical algorithm for solving the order placement problem in a general case and our simulation results, and Section \ref{sec.conc} concludes. \section{The order placement problem}\label{sec.routing} Consider a trader who has a mandate to buy $S$ shares of a stock within a (short) time interval $[0,T]$. The deadline $T$ may be a fixed time horizon (e.g. 1 minute) or a stopping time (e.g. triggered by price changes or trading volume dynamics). To gain queue priority, the trader can immediately submit $K$ limit orders of sizes $L_k$ to multiple exchanges $k=1,\dots,K$ and also market orders for $M$ shares. The trader's {\it order placement} decision is thus summarized by a vector $X~\overset{\Delta}{=}~(M,L_1,\dots,L_K)\in\mathbb{R}^{K+1}_+$ of order sizes whose components are non-negative (only buy orders are allowed). Our objective is to define a meaningful framework in which the trader may choose between various possibilities for this order placement decision, for example between sending an order to a single exchange or splitting it in some proportion across $K$ exchanges. Our focus here is on limit order placement and we assume for simplicity that a market order of any size up to $S$ can be filled immediately and with certainty at any exchange. Under this assumption sending market orders to exchanges with high fees is clearly sub-optimal and we therefore consider a single exchange with the smallest fee $f$ for the purpose of sending a single market order of size $M$. This simplifying assumption is reasonable as long as the target size $S$ is of the same order of magnitude as the prevailing market depth. Otherwise the quantity $S$ can be filled with multiple market orders at exchanges with progressively larger fees $f_1<f_2<\dots<f_K$, and the total cost of these market orders becomes a convex piecewise linear function. Our results extend to this case, but to avoid additional notation we assume that $S$ can be executed with a single market order. Limit orders with quantities $(L_1,\dots,L_K)$ join queues of $(Q_1,\dots,Q_K)$ pre-existing orders at the best bids of $K$ limit order books, where $Q_k\geq 0$. Empty queues are allowed which corresponds to sending an order inside the bid-ask spread. As a simplification we assume that all $K$ limit orders submitted by the trader have the same price (the national best bid price), but limit orders submissions deeper in the order book can also be treated in our formulation at the expense of additional notation. Denote by $(x)_+\overset{\Delta}{=}\max(x,0)$. If the trader does not modify his limit orders before time $T$, the amount purchased with each limit order by time $T$ can be explicitly computed as a function of a future order flow: $$\min(\max(\xi_k-Q_k,0),L_k)=(\xi_k-Q_k)_+-(\xi_k-Q_k-L_k)_+,\quad k=1,\dots,K$$ \noindent where $\xi_k$ is a total outflow from the front of the $k$-th order queue. The order outflow $\xi_k$ consists of order cancelations that occurred before time $T$ from queue positions in front of an order $L_k$, and of marketable orders that reach the $k$-th exchange before $T$. The mechanics of limit order fills and order outflows are further illustrated on Figure \ref{fig.queue_fill}. \begin{figure}[ht] \noindent \begin{center} { \includegraphics[width=60mm]{queue_fill} } \end{center} \caption{Limit order execution on exchange $k$ depends on the order size $L_k$, the queue $Q_k$ in front of it, total sizes of order cancelations $C_k$ and marketable orders $D_k$, specifically on $\xi_k=C_k+D_k$.} \label{fig.queue_fill} \end{figure} We note that limit order fill amounts are random because they depend on random future bid queue outflows $\xi=(\xi_1,\dots,\xi_K)$. The random variable $\xi$ is defined with respect to an execution time horizon $T$, therefore its distribution depends on a trader's choice of $T$. Here we do not make any assumptions regarding the distribution of $\xi$ except for illustration purposes. In fact our formulation leads to an intuitive iterative procedure that approximates the optimal order allocation by using historical data and also does not require specifying a distribution for $\xi$. \newpage The {\it total amount} of shares $A(X,\xi)$ bought by the trader by time $T$ with his limit and market orders is a function of the order allocation $X$ and an overall bid queue outflow $\xi$: \begin{equation}\label{optim.amt} A(X,\xi)=M+\sum_{k=1}^K\left((\xi_k-Q_k)_+-(\xi_k-Q_k-L_k)_+\right) \end{equation} The total price of this purchase is divided into a benchmark cost paid regardless of trader's decisions, computed using a mid-quote price level, and an {\it execution cost} relative to mid-quote price given by \begin{equation}\label{optim.cost} (h+f)M-\sum_{k=1}^K(h+r_k)((\xi_k-Q_k)_+-(\xi_k-Q_k-L_k)_+), \end{equation} \noindent where $h$ is a half of the bid-ask spread at time 0, $f$ is the lowest available fee for taking liquidity and $r_k, k=1,\dots,K$ are rebates for adding liquidity on different exchanges. Most exchanges in the U.S. pay rebates as an incentive for traders to send passive limit orders, and charge fees for executing marketable orders. In contrast, {\it inverse} exchanges pay traders for executing marketable orders and charge for executing passive limit orders. For example, at the time of writing a U.S. equity exchange Direct Edge EDGA had a negative rebate $r=-\$0.0006$ per share for passive orders and a negative fee $f=-\$0.0004$ per share for marketabe orders. Another inverse exchange BATS BYX had a rebate $r=-\$0.0002$ and a fee $f=-\$0.0002$. In addition to fees and rebates, the parameters $f,~r_k,~k=1,\dots,K$ can include other explicit or implicit trading costs, such as exchange messaging fees or average adverse selection for executed limit orders (see e.g. \cite{Moallemi2011}). The benchmark price in our formulation is the mid-quote price at time 0, so in \eqref{optim.cost} the trader saves half of the bid-ask spread plus liquidity rebates on his limit orders, and pays half of the spread plus a liquidity fee on his market orders. Limit orders reduce the cost but lead to a risk of falling behind the target quantity $S$ because their fills are random. To capture this {\it execution risk} we include, in the objective function, a penalty for violations of target quantity in both directions: \begin{equation}\label{optim.tarpen} \lambda_u\left(S - A(X,\xi)\right)_+ +\lambda_o\left(A(X,\xi)-S\right)_+ , \end{equation} \noindent where $\lambda_u\geq 0,\lambda_o\geq 0$ are marginal penalties in dollars per share for, respectively falling behind or exceeding the execution target $S$. These penalties are motivated by a correlation that exists between limit order executions and price movements (so-called adverse selection, see e.g. \cite{sandaas2001}). If $A(X,\xi)<S$, the trader has to purchase the remaining $S-A(X,\xi)$ shares at time $T$ with market orders. Adverse selection implies that conditionally on the event $\{A(X,\xi)<S\}$ prices have likely moved up and the transaction cost of market orders at time $T$ is higher than their cost at time 0, i.e. $\lambda_u>h+f$. Alternatively, if $A(X,\xi)>S$ the trader experiences buyer's remorse - conditionally on this event prices have likely moved down and he could have achieved a better execution by sparing some of his market orders. Besides adverse selection, parameters $\lambda_u,\lambda_o$ may reflect trader's execution preferences. For example a trader with a positive forecast of short-term returns may prefer to trade early with a market order and set a larger value for $\lambda_u$ compared to $\lambda_o$. A trader with significant aversion to execution risk would choose high values for both $\lambda_u,\lambda_o$ to reflect this aversion. \begin{myproblem}[Optimal order placement problem] An {\it optimal order placement} is a vector $X^*\in\mathbb{R}^{K+1}_+$ solution of \begin{equation}\label{optim.set} \min_{X\in\mathbb{R}^{K+1}_+}\mathbb{E}[ v(X,\xi) ] \end{equation} where \begin{equation}\label{eq.V} v(X,\xi):=(h+f)M-\sum_{k=1}^K(h+r_k)((\xi_k-Q_k)_+-(\xi_k-Q_k-L_k)_+)+ \lambda_u\left(S - A(X,\xi))\right)_+ +\lambda_o\left(A(X,\xi)-S\right)_+ \end{equation} is the sum of the execution cost and penalty for execution risk. \end{myproblem} We will denote $V(X)=E[ v(X,\xi)]$. The use of expected value in the objective function is motivated by the fact that short-term order placement decisions are made by trading algorithms a large number of times each day, and assuming that the order flow over the trading horizon is described by an ergodic process, average costs converge to their expected value over time. We begin by assuming certain economically reasonable restrictions on parameter values. \noindent {\bf Assumptions} \noindent {\bf A1} $\underset{k}{\min}\{r_k\}+h > 0$: even if some rebates $r_k$ are negative, limit orders reduce the execution cost. \noindent {\bf A2} $\lambda_o > h+\underset{k}{\max}\{r_k\}$ and $\lambda_o > -(h+f)$: there is no incentive to exceed the target quantity $S$ regardless of fees and rebates, even if they are negative. \noindent Although negative rebate values are possible, in the U.S. they are typically smaller than the smallest possible value of $h=\$0.005$, justifying our assumptions A1-A2. Proposition \ref{prop.lubound} below shows that it is not optimal to submit limit or market orders that are {\it a priori} too large or too small (larger than the target size $S$ or whose sum is less than $S$). Proposition \ref{prop.objfun} guarantees the existence of an optimal solution. \begin{myprop}\label{prop.lubound} Consider the compact convex subset of $\mathbb{R}^{K+1}_+$ defined by $$\mathcal{C}\overset{\Delta}{=}\left\{X\in\mathbb{R}^{K+1}_+\Bigm| 0\leq M\leq S,~~0\leq L_k\leq S-M, k=1,\dots,K, ~~M+\sum_{k=1}^K{L_k}\geq S\right\}$$ \noindent Under assumptions A1-A2 for any $\tilde X\notin\mathcal{C}$, $\exists \tilde X'\in\mathcal{C}$ with $V(\tilde X')\leq V(\tilde X)$. Moreover, if $\underset{k}{\min}\left\{\mathbb{P}(\xi_k>Q_k+S)\right\}>0$, the inequality is strict: $V(\tilde X')<V(\tilde X)$. \end{myprop} Proposition \ref{prop.lubound} shows that it is never optimal to overflow the target size $S$ with a single order, but it may be optimal to exceed the target $S$ with the sum of order sizes $M+\sum_{k=1}^K{L_k}$. The penalty function (\ref{optim.tarpen}) effectively implements a soft constraint for order sizes and focuses the search for an optimal order allocation to the set $\mathcal{C}$. Specific economic or operational considerations could also motivate adding hard constraints to problem \eqref{optim.set}, e.g. $M=0$ or $\sum_{k=1}^K{L_k}=S$. Such constraints can be easily included in our framework but absent the aforementioned considerations we do not impose them here. \begin{myprop}\label{prop.objfun} Under assumptions A1-A2, $V:\mathbb{R}^{K+1}_+\mapsto \mathbb{R}$ is convex, bounded from below and has a global minimize $~X^\star\in\mathcal{C}$. \end{myprop} We may also consider an alternative approach to order placement optimization, which turns out to be related to our original formulation by duality. Consider the following problem: \begin{myproblem}[Alternative formulation: cost minimization under execution constraints] \begin{align} &\min_{X\in\mathbb{R}^{K+1}_+}\mathbb{E}[(h+f)M-\sum_{k=1}^K(h+r_k)((\xi_k-Q_k)_+-(\xi_k-Q_k-L_k)_+)]\label{optim.set.constr}\\ &\text{subject to:}\quad\mathbb{E}\left[\left(S - A(X,\xi)\right)_+\right]\leq\mu_u \label{optim.constr.1}\\ &\hskip 60pt \mathbb{E}\left[\left(A(X,\xi) - S\right)_+\right]\leq\mu_o \label{optim.constr.2} \end{align} \end{myproblem} \noindent In this alternative formulation a trader can specify his tolerance to execution risks using constraints on expected order shortfalls and overflows. The goal is to minimize an expectation of order execution costs under the expected shortfall constraints. The Problem 2 does not appear to be tractable, but it has a convex objective and convex inequality constraints, and we can easily find its (Lagrangian) dual problem: \begin{myproblem} \begin{equation}\label{optim.set.dual} \max_{\lambda_u\geq 0,\lambda_o\geq 0}\left\{V^\star(\lambda_u,\lambda_o)-\lambda_u\mu_u-\lambda_o\mu_o\right\} \end{equation} where $V^\star(\lambda_u,\lambda_o) = \underset{X\in\mathbb{R}^{K+1}_+}{\min}\mathbb{E}[v(X,\xi)]$ is the optimal objective value from Problem 1 given parameter values $\lambda_u,\lambda_o$. \end{myproblem} \noindent We see that the Problem 3 is related to our original order placement problem - solving the Problem 3 (and therefore, the Problem 2) amounts to re-solving the Problem 1 for different values of $\lambda_u,\lambda_o$. This discussion also leads to a new interpretation of parameters $\lambda_u,\lambda_o$ in the Problem 1 as shadow prices for expected shortfall and overflow constraints in the related Problem 2. Hereafter we focus on the (more tractable) Problem 1, but note that the optimal point for the Problem 2 can also be found by solving its dual problem. \section{Optimal order allocation} \subsection{Choice of order type: limit orders vs market orders}\label{sec.1exchg} To highlight the tradeoff between limit and market order executions in our optimization setup, we first consider a case when the asset is traded on a single exchange, and the trader has to choose an optimal split between limit and market orders. Since $K=1$, we suppress the subscript 1 throughout this section. \begin{myprop}[Single exchange: optimal split between limit and market orders]\label{prop.1exchg} Assume that $\xi$ has a continuous distribution and (A1-A2) hold. The optimal order allocation depends on $\lambda_u$: If $\lambda_u\leq\underline{\lambda_u}\overset{\Delta}{=}\dfrac{2h+f+r}{F(Q+ S)}-(h+r)$, the optimal strategy is to submit only limit orders: $\quad(M^{\star},L^{\star})=(0,S)$. If $\lambda_u\geq\overline{\lambda_u}\overset{\Delta}{=}\dfrac{2h+f+r}{F(Q)}-(h+r)$, the optimal strategy is to submit only market orders: $\quad(M^{\star},L^{\star})=(S,0)$. If $\lambda_u\in(\underline{\lambda_u},\overline{\lambda_u})$, the optimal split between limit and market orders is \begin{equation}\label{sol.1exchg} \left\{ \begin{aligned} & M^{\star}=S - F^{-1}\left(\frac{2h+f+r}{\lambda_u+h+r}\right)+Q, \\ & L^{\star}=F^{-1}\left(\frac{2h+f+r}{\lambda_u+h+r}\right)-Q, \end{aligned} \right. \end{equation} \noindent where $F(\cdot)$ is a cumulative distribution function of the bid queue outflow $\xi$. \end{myprop} \noindent In the case of a single exchange, Proposition \ref{prop.lubound} implies that $M^{\star}+L^{\star}=S$, therefore there is no risk of exceeding the target size and $\lambda_o$ does not affect the optimal solution. The trader is only concerned with the risk of falling behind the target quantity, and balances this risk with the fee, rebate and other market information. The parameter $\lambda_u$ can be interpreted as trader's urgency to fill the orders, and higher values of $\lambda_u$ lead to smaller limit order sizes, as illustrated on Figure \ref{fig.Sl_onex}. In contrast, the optimal market order size increases with $\lambda_u$. The optimal split between market and limit orders depends on the ratio $\frac{2h+f+r}{\lambda_u+(h+r)}$ which balances marginal costs and savings from a market order. It also depends on the order outflow distribution $F(\cdot)$ and the queue length $Q$ - keeping all else constant, a trader would submit a larger limit order if its execution is more likely and vice versa. The optimal limit order size decreases with $\lambda_u$ as it becomes more expensive to underfulfill the order and increases with $f$ as market orders become more expensive. Another interesting feature is that $L^{\star}$ is fully determined by $Q$, $F$ and pricing parameters $h,r,f,\lambda_u$, while $M^{\star}$ increases with $S$. The consequence of this solution feature is that as the target size $S$ increases, a larger fraction $\frac{M^{\star}}{S}$ of it is executed with a market order. The total quantity that can realistically be filled with a limit order is limited by $Q$ and $\xi$, so to accommodate larger target sizes the trader resorts to market orders. This {\it bounded capacity} feature of limit orders also appears in our solutions for multiple exchanges. For example, as the number of available exchanges $K$ increases, the overall prospects of filling limit orders at any of them improve and the fraction $\frac{M^{\star}}{S}$ decreases. We can see that the solution $(M^\star, L^\star)$ depends on the entire distribution $F(\cdot)$ and not just on the mean of $\xi$. Limit orders are filled when $\xi\geq Q+L$, so the tail of $F(\cdot)$ affects order executions and is an important determinant of the optimal order allocation. Figure \ref{fig.Sl_onex} shows two order allocations for exponential and Pareto distributions of $\xi$ with equal means. \begin{figure}[h!] \noindent \begin{center} { \includegraphics[width=100mm]{Sl_onex} } \end{center} \caption{Optimal limit order size $L^{\star}$ for one exchange. The parameters for this figure are: $Q = 2000,S = 1000, h = 0.02, r = 0.002, f = 0.003$. Colors correspond to different order outflow distributions - exponential with means 2200 and 2500 and Pareto with mean 2200 and a tail index 5.} \label{fig.Sl_onex} \end{figure} \subsection{Optimal routing of limit orders across multiple exchanges}\label{sec.2exchg} When multiple trading venues are available, dividing the target quantity among them reduces the risk of not filling the order and may improve the execution quality. However, sending too many orders leads to an undesirable possibility of exceeding the target size. Proposition \ref{prop.foc} gives optimality conditions for an order allocation $X^{\star}=(M^{\star},L^{\star}_1,\dots,L^{\star}_K)$ that balances shortfall risks and costs. The following probabilities play an important role in this balance: $$p_0\overset{\Delta}{=}\mathbb{P}\left(\underset{k=1}{\overset{K}{\bigcap}}\{\xi_k\leq Q_k\}\right), \quad p_j\overset{\Delta}{=}\mathbb{P}\left(\underset{k\neq j}{\bigcap}\{\xi_k\leq Q_k\}\bigg|\xi_j>Q_j\right),\quad j=1,\dots,K$$ Intuitively, $p_0$ is a probability that no limit orders will be filled given current queue sizes, and it measures the overall execution prospects for limit orders. Each $p_j$ is a probability of no fills everywhere except the $j$-th exchange, conditional on a fill at the exchange $j$, so $p_j$ measure tail dependences between order flows on different exchanges. \begin{myprop}\label{prop.foc} Assume (A1-A2), also assume that the distribution of $\xi$ is continuous, $\underset{k}{\max}\left\{F_k(Q_k+S)\right\}<1$ and $\lambda_u<\underset{k}{\max}\left\{\dfrac{2h+f+r_k}{F_k(Q_k)}-(h+r_k)\right\}$. Then: \vskip 6pt \begin{enumerate} \item If \begin{equation}\label{cond.mrkord} \lambda_u\geq\frac{2h+f+\underset{k}{\max}\{r_k\}}{p_0}-(h+\underset{k}{\max}\{r_k\}) \end{equation} then any optimal order placement strategy involves market orders: $M^\star>0$. \item If \begin{equation}\label{cond.limord} p_j>\frac{\lambda_o-(h+r_j)}{\lambda_u+\lambda_o} \end{equation} then any optimal order placement strategy involves limit orders on the $j$-th exchange: $L_j^\star>0$. \item If \eqref{cond.mrkord}-\eqref{cond.limord} hold for all exchanges $j=1,\dots,K$, then $X^\star\in\mathcal{C}$ is an optimal order placement if and only if the following conditions are fulfilled: \begin{align} \mathbb{P}\left(M^{\star}+\sum\limits_{k=1}^K\left((\xi_k-Q_k)_+-(\xi_k-Q_k-L^{\star}_k)_+\right)<S\right)=\frac{h+f+\lambda_o}{\lambda_u+\lambda_o}& \label{FOC1}\\ \mathbb{P}\left(M^{\star}+\sum\limits_{k=1}^K\left((\xi_k-Q_k)_+-(\xi_k-Q_k-L^{\star}_k)_+\right)<S \bigg| \xi_j>Q_j+L^{\star}_j\right)&=\frac{\lambda_o-(h+r_j)}{\lambda_u+\lambda_o}, \notag\\ &j=1,\dots, K \label{FOC2} \end{align} \end{enumerate} \end{myprop} Equations \eqref{FOC1}-\eqref{FOC2} show that an optimal order allocation equates shortfall probabilities to specific values computed with pricing parameters. This gives yet another interpretation for parameters $\lambda_u,\lambda_o$ - a trader can specify his tolerance for execution risk in terms of shortfall probabilities and use the above equations to calibrate these parameters. When the number of exchanges $K$ is large, the probabilities in \eqref{FOC1}-\eqref{FOC2} are difficult to compute in closed-form. However, the case $K=2$ is relatively tractable and will be analyzed as an illustration. The assumption of independence between $\xi_1,\xi_2$ is made only in this example and is not required for the rest of our results. In Section \ref{sec.sa} we study the effect of correlation between order flows on optimal order placement decisions. \newpage \noindent {\bf Corollary} {\it Consider the case of two exchanges with outflows $\xi_1,\xi_2$ that are independent and have continuous distributions. If \begin{enumerate} \item $\underset{k=1,2}{\max}\left\{F_k(Q_k+S)\right\}<1$, \item $\lambda_u<\underset{k=1,2}{\max}\left\{\dfrac{2h+f+r_k}{F_k(Q_k)}-(h+r_k)\right\}$, $\lambda_u\geq\dfrac{2h+f+\underset{k=1,2}{\max}\{r_k\}}{F_1(Q_1)F_2(Q_2)}-(h+\underset{k=1,2}{\max}\{r_k\})$, and \vskip 6pt \item $F_1(Q_1)<1-\dfrac{h+r_2}{\lambda_o}, F_2(Q_2)<1-\dfrac{h+r_1}{\lambda_o}$ \vskip 6pt \end{enumerate} then there exists an optimal order allocation $X^\star=(M^{\star}, L^{\star}_1, L^{\star}_2)\in int\{\mathcal{C}\}$ and it verifies \small \begin{subequations} \begin{align} & L^{\star}_1 = Q_2 + S - M^{\star} - F^{-1}_2\left(\frac{\lambda_o-(h+r_1)}{\lambda_u+\lambda_o}\right) \label{sol.2exchg.1}\\ & L^{\star}_2 = Q_1 + S - M^{\star} - F^{-1}_1\left(\frac{\lambda_o-(h+r_2)}{\lambda_u+\lambda_o}\right) \label{sol.2exchg.2}\\ & \bar F_1( Q_1+L^{\star}_1)\bar F_2(Q_2+S- M^{\star}-L^{\star}_1) +\int\limits_{Q_1+ S-M^{\star}-L^{\star}_2}^{Q_1+L^{\star}_1}{\bar F_2(Q_1+Q_2+S-M^{\star}-x_1)dF_1(x_1)}= \frac{\lambda_u-(h+f)}{\lambda_u+\lambda_o}\label{sol.2exchg.m}, \end{align} \end{subequations} \normalsize \noindent where $F_1(\cdot),F_2(\cdot)$ are the cdf of $\xi_1,\xi_2$ respectively. } \noindent In the solution \eqref{sol.2exchg.1}--\eqref{sol.2exchg.2} optimal limit order quantities $L^{\star}_1,L^{\star}_2$ are linear functions of an optimal market order quantity $M^{\star}$. When \eqref{sol.2exchg.1}-\eqref{sol.2exchg.2} are substituted into \eqref{sol.2exchg.m} we obtain a (non-linear) equation for $M^{\star}$, which can be solved for a given distribution of $(\xi_1,\xi_2)$. \iffalse \noindent{\bf Example} If $\xi_1, \xi_2$ are exponentially distributed with means $\mu_1,\mu_2$ respectively, then an optimal order allocation is given by: \begin{equation}\label{sol.2exchg.exp_lim} \left\{ \begin{aligned} &M^{\star} = Q_1+Q_2+S - z\\ &L^{\star}_1 = z - Q_1 + \mu_2\log\left(\frac{\lambda_u+h+r_1}{\lambda_u+\lambda_o}\right) \\ &L^{\star}_2 = z - Q_2 + \mu_1\log\left(\frac{\lambda_u+h+r_2}{\lambda_u+\lambda_o}\right), \end{aligned} \right. \end{equation} \noindent where $z$ is a solution of a transcendental equation: \begin{align} &1+\log\left(\frac{(\lambda_u+h+r_1)(\lambda_u+h+r_2)}{(\lambda_u+\lambda_o)^2}\right)+\frac{z}{\mu} =\frac{\lambda_u-(h+f)}{\lambda_u+\lambda_o}e^{\frac{z}{\mu}}~\text{, if}~\mu_1=\mu_2=\mu \label{sol.2exchg.exp_mrk_eq} \\ &\text{or}\notag\\ &\frac{\mu_1}{\mu_1-\mu_2}e^{-\frac{z}{\mu_1}}\left(\frac{\lambda_u+h+r_1}{\lambda_u+\lambda_o}\right)^{\frac{\mu_1-\mu_2}{\mu_1}} +\frac{\mu_2}{\mu_2-\mu_1}e^{-\frac{z}{\mu_2}}\left(\frac{\lambda_u+h+r_2}{\lambda_u+\lambda_o}\right)^{\frac{\mu_2-\mu_1}{\mu_2}} = \frac{\lambda_u-(h+f)}{\lambda_u+\lambda_o}~\text{, if}~\mu_1\neq\mu_2 \label{sol.2exchg.exp_mrk_neq} \end{align} \noindent Similarly to the case of one exchange, in this example an optimal market order size $M^\star$ is an increasing linear function of queue sizes $Q_1, Q_2$ and the target quantity $S$, while optimal limit order sizes $L^\star_i$ are decreasing functions of the corresponding queue sizes $Q_i$. As in the case of one exchange, the optimal limit order sizes do not depend on the target quantity, but the optimal market order size increases with it. In addition we note that each $L^\star_i$ depends on the order flow distribution on both exchanges through $\mu_{1,2}$ and $z$. \fi \section{An algorithm for optimal order placement}\label{sec.sa} \subsection{A stochastic algorithm based on (re)sampling of recent order flow} Practical applications require a fast and flexible method for optimizing order placement across multiple trading venues. As their number $K$ increases, it becomes progressively more difficult to calculate the objective function $V(X)$ in Problem 1 which is a $K$-dimensional integral, and also to obtain analytical solutions for that problem. However, numerical solutions can be efficiently computed even in high dimensions based on stochastic approximation methods. These methods use samples of the order outflows $\xi_k$ and approximate the gradient of $V(X)$ along a random optimization path. Applying this approach for our problem formulation yields an intuitive iterative algorithm that updates trader's order allocation in response to past order execution outcomes. \newpage Our numerical solution is based on the robust stochastic approximation algorithm of \cite{Nemirovski2009}. Consider an objective function $V(X)\overset{\Delta}{=} \mathbb{E}[v(X,\xi)]$ to be minimized and denote by $g(X,\xi)\overset{\Delta}{=}\nabla v(X,\xi)$ where the gradient is taken with respect to $X$. The stochastic approximation algorithm tackles the problem of minimizing $V(X)$ in the following way: \vskip 6pt \begin{algorithmic}[1] \STATE Choose $X_0\in\mathbb{R}^{K+1}$ and a step size $\gamma$; \FOR{$n=1,\dots,N$} \STATE Sample a random variable $\xi^n\in\mathbb{R}^{K}$ from its distribution $F$ \STATE Set an iterate $X_{n}=X_{n-1}-\gamma g(X_{n-1},\xi^n)$ \ENDFOR \STATE \textbf{end} \STATE Set the numerical solution $\hat X^\star\overset{\Delta}{=}\frac{1}{N}\sum_{n=1}^NX_n$ \end{algorithmic} \vskip 6pt \noindent Here random variables $\xi^n$ are independent across iterations, but the $K$ components of each draw $\xi^n$ need not be independent. Under some weak assumptions, satisfied by our objective function, the numerical solution $\hat X^\star$ converges to the optimal solution $X^\star$ and has a performance bound \newline $V(\hat X^\star)-V(X^\star)\leq\frac{DM}{\sqrt{N}}$, where $D=\underset{X,X'\in\mathcal{C}}{\max}\|X-X'\|_2$, $M=\sqrt{\underset{X\in\mathcal{C}}{\max}~\mathbb{E}\left[\|g(X,\xi)\|^2_2\right]}$. \noindent The optimal step size is $\gamma^\star=\frac{D}{\sqrt{N}M}$ and we use a step size $$\gamma=K^{1/2}S\left(N(h+f+\lambda_u+\lambda_o)^2+N\sum\limits_{k=1}^K(h+r_k+\lambda_u+\lambda_o)^2\right)^{-1/2}$$ \noindent which scales appropriately with problem parameters. For more details on stochastic approximation methods we refer to \cite{Kushner2003} and \cite{Nemirovski2009}. In general, this method requires drawing $K$-dimensional random variables $\xi$ from a possibly complicated distribution $F$, but in our case one may avoid specifying $F$ because the function $g(X,\xi)$ takes a particular form: \small $$ g(X_n,\xi)= \left( \begin{matrix} & h + f -\lambda_u\indicator{A(X_n,\xi)<S} + \lambda_o\indicator{A(X_n,\xi)>S}\\ \\ &\bigl(-(h+r_1)-\lambda_u\indicator{A(X_n,\xi)<S} + \lambda_o\indicator{A(X_n,\xi)>S}\bigr)\indicator{\xi_1>Q_1+L_{1,n}} \\ &\dots\\ &\bigl(-(h+r_K) -\lambda_u\indicator{A(X_n,\xi)<S} + \lambda_o\indicator{A(X_n,\xi)>S}\bigr)\indicator{\xi_K>Q_K+L_{K,n}} \\ \end{matrix} \right) $$ \normalsize \vskip 6pt \noindent We note that $g(X_n,\xi)$ depends on random variables $\xi$ only through indicator functions, which have specific economic meaning: $\indicator{A(X_n,\xi)<S}=1$ if on the $n-$th iteration the trader fell behind the target quantity, $\indicator{A(X_n,\xi)>S}=1$ if the target was exceeded, and $\indicator{\xi_k>Q_k+L_{k,n}}=1$ if a limit order on exchange $k$ was fully executed. Having recorded these indicators for past executions, a trader can resample them later according to this method and optimize future order allocations. Alternatively, one can also compute these indicator functions based on recent order flow data. Our iterative solution updates order sizes in response to previous order execution outcomes in an intuitive way. For example, the first component of $g(X_n,\xi)$ describes updates of the market order size $M$ - on each iteration $M$ is decreased by $\gamma(h + f)$ to reduce a cost of execution, but if a trader fell behind his target quantity, $M$ will be increased on the next step by $\gamma\lambda_u$ to reduce execution risk. Since overtrading is also penalized, $M$ is decreased by $\gamma\lambda_o$ whenever a target quantity is exceeded. Limit order sizes are updated similarly, but only when these orders are executed. This iterative algorithm gives a specific way to resample past order fill data or historical order flow data to obtain a solution for the order placement problem. The algorithm consists of sequential updates and involves only basic arithmetic operations, so it can be implemented for on-line order routing optimization in real-time trading applications. Alternatively, one may follow the same steps using historical data or an order flow model to generate samples of $\xi$ and pre-compute optimal order allocations off-line. To illustrate our method, we simulated its application using historical data. We considered an execution of a moderate-sized order to buy $S=2000$ shares of Microsoft Corporation (MSFT) stock with an execution deadline $T=1$ minute, to be executed on the NASDAQ and BATS Z stock exchanges. This liquid stock is traded on multiple exchanges, but for simplicity we considered only these two. We assumed in the simulation that rebates are $r_{NSDQ}=20$ and $r_{BATS}=25$ mills (1 mill = $10^{-4}$ dollars) per share which is close to their historical averages. The fee was assigned to $f=29$ mills per share and the half-spread was set to $h=50$ mills which is typical for this stock. To perform numerical optimization we used trade and quote (TAQ) data from January to March 2012, and computed one-minute seller-initiated trade volumes on these two exchanges using the \cite{lee91} rule with a zero millisecond time offset between trades and quotes. The seller-initiated trade volumes were used as estimates of one-minute order outflows $\xi_{BATS},\xi_{NSDQ}$ from the corresponding best bid queues. These estimates are conservative because they do not include possible order cancelations from the front of the queues. TAQ data does not allow to directly track orders and order cancelations, but this information is available in more detailed ``Level 2" datasets. To ensure that seller-initiated trading volume corresponds to order outflows from the same queue we only considered one-minute samples where the best bid price did not change. Although it was possible to compute a separate optimal solution for each configuration of the problem parameters (initial queue sizes $Q_{BATS},Q_{NSDQ}$, target size $S$, etc.), we decided to reduce the overall number of distinct solutions by using the same order allocation for similar queue size configurations. To estimate these solutions, we separated historical data into terciles (three equal-sized subsamples), sorting the data by $Q_{BATS},Q_{NSDQ}$, and by trading volume in the previous minute $VOL_{BATS},VOL_{NSDQ}$, from low to high values. This created $81$ data subsamples (3 per parameter) and we estimated a separate numerical solution for each subsample. In the resulting solution table each order allocation takes into account the magnitude of $Q_{BATS},Q_{NSDQ}$ and $VOL_{BATS},VOL_{NSDQ}$, up to a tercile. Table 1 presents a subset of these order allocations for different levels of $VOL_{BATS}$ and for $Q_{BATS},Q_{NSDQ},VOL_{NSDQ}$ fixed in their corresponding low terciles. \begin{table}[!h] \begin{center} \begin{tabular}{|c|c|c c c|} \hline $\lambda_u,\lambda_o$ & $VOL_{BATS}$ & $\big(M^\star$ & $L^\star_{BATS}$ & $L^\star_{NSDQ}\big)/S$\\ \hline \multirow{3}{*}{80 mills} & low & 0\% & 100\% & 100\%\\ & medium & 0\% & 100\% & 100\%\\ & high & 0\% & 100\% & 28\%\\ \hline \multirow{3}{*}{120 mills} & low & 99\% & 1\% & 1\%\\ & medium & 98\% & 2\% & 1\%\\ & high & 9\% & 85\% & 15\%\\ \hline \end{tabular} \caption{Examples of optimal order allocation on BATS (Z) and \newline NASDAQ for buying 2000 shares of MSFT.} \end{center} \end{table} This example demonstrates how an optimal order allocation shifts from limit orders to market orders as a trader's aversion to execution risk (represented by $\lambda_u,\lambda_o$) increases. We also notice that an order allocation depends on $VOL_{BATS}$ - since trading volume is positively autocorrelated in time, a ``high" reading of volume in the previous minute predicts a higher volume in the next minute, making a limit order execution more likely on the BATS Z exchange. Therefore the numerical solution places more limit orders there. After computing the numerical solutions, we compared their average costs and the costs for an equal split strategy $X_E = \left(\frac{S}{3},\frac{S}{3},\frac{S}{3}\right)$, using an additional month of data from April 2012. For an $n-$th test sample we selected an appropriate numerical solution from our table and computed $v(\hat X^\star,\xi_n)$ for that sample. The average cost per share for $\hat X^\star$ is $W(\hat X^\star) = \frac{1}{NS}\sum_{n=1}^{N}v(\hat X^\star,\xi_n)$, where $N$ is the number of test samples, and the expression for $W(X_E)$ is similar. Overall we found that order placement optimization can substantially reduce the costs of trading. When risk aversion parameters are set to $\lambda_u=\lambda_o=80$ mills, the execution risk is unimportant relatively to the execution cost. In this case the average execution amount with our strategy $\hat X^\star$ is 533 shares compared to 867 shares with $X_E$. However, $\hat X^\star$ achieves a much lower cost with limit orders. The average cost per share with the allocation $\hat X^\star$ is $50.21$ basis points, including the penalty for execution shortfall. This is 22\% lower than the average cost obtained with $X_E$, $64.37$ basis points per share. When we set $\lambda_u=\lambda_o=120$ mills, the execution shortfall and overtrading penalties contributed more to the objective function. While the benchmark $X_E$ remained the same, the optimal solution $\hat X^\star$ shifted to market orders attempting to reduce the execution risk and achieving an average execution quantity of 1763 shares, compared to 867 shares for $X_E$. The average cost per share is now 80.21 basis points with $\hat X^\star$, 8\% lower than the benchmark cost of 87.03 basis points for $X_E$. \subsection{Market fragmentation and costs of trading}\label{ssec.fragm} \noindent In a fragmented financial market high-frequency traders and other trading algorithms simultaneously act on multiple trading venues, connecting them (see e.g. \cite{Menkveld2011}) and creating positive correlations between order flows. To investigate how these correlations affect order placement strategies and order execution costs we studied an example with two exchanges where order flows were generated by a factor model: $$\xi_1 = \alpha\xi_0+(1-\alpha)\epsilon_1,\quad\xi_2 = \alpha\xi_0+(1-\alpha)\epsilon_2,$$ \noindent The factors $\xi_0,\epsilon_1,\epsilon_2\sim Pois(\mu T)$ are i.i.d. random variables and the scalar parameter $\alpha\in[0,1]$ controls the degree of positive correlation between order flows $\xi_1$ and $\xi_2$. For the purpose of this example we assumed that a trader is buying $S=1000$ shares of a stock with a deadline $T=1$ minute, initial queue sizes $(Q_1,Q_2) = (1900,2000)$ shares and $\mu=2200$ shares per minute. The pricing parameters are set to typical values for U.S. equities: $h=200$, $r=20$, $f=30$, and $\lambda_o=240$, $\lambda_u=260$ mills per share. In the numerical solutions, plotted on Figure \ref{fig.twoex_fig_corr}, we can observe that as the parameter $\alpha$ increases, the sum of order sizes $M+L_1+L_2$ decreases to the target quantity $S=1000$. When $\alpha\approx 1$, the second exchange does not provide any benefit to a trader in terms of diversifying his execution risk, so it is not optimal anymore to oversize the total quantity of limit orders. This case is similar to the case of a single exchange where the total order quantity $M+L=S$ by Proposition \ref{prop.1exchg}. Executing orders on two exchanges with positively correlated order flows $(\xi_1,\xi_2)$ and queue sizes $(Q_1,Q_2)$ is not equivalent to having a single consolidated exchange with $\xi=\xi_1+\xi_2$ and $Q=Q_1+Q_2$, even if the correlation between $\xi_1$ and $\xi_2$ is high. Figure \ref{fig.twoex_fig_corr_cost} compares average execution costs for these two cases with an optimal order allocation and different values $\alpha$. It appears from this example that the availability of multiple exchanges reduces the costs of trading compared to a single consolidated exchange, but only as long as multiple exchanges remain relatively uncorrelated. \newpage \begin{figure}[h!] \noindent \begin{center} { \includegraphics[width=150mm]{twoex_fig_corr} } \end{center} \caption{Optimal order allocations with correlated order flows. Left axis - order sizes $(M,L_1,L_2)$ from the numerical solution. Right axis - their sum $M+L_1+L_2$. } \label{fig.twoex_fig_corr} \end{figure} \begin{figure}[h!] \noindent \begin{center} { \includegraphics[width=150mm]{twoex_fig_corr_cost} } \end{center} \caption{Comparison of average order execution costs: two exchanges with correlated order flows against a single consolidated exchange. Dashed lines show 95\% confidence intervals for the averages.} \label{fig.twoex_fig_corr_cost} \end{figure} \newpage \subsection{Convergence of numerical solutions and sensitivity analysis} We also applied the algorithm to several numerical examples to assess its convergence and the sensitivities of a numerical solution to various problem inputs. First, we considered an example where a trader needs to buy $S=1000$ shares of a stock with a deadline $T=1$ minute, using a single exchange with an initial bid queue size $Q=2000$ shares, order outflow $\xi\sim Pois(\mu T)$ and $\mu=2200$ shares per minute. With these parameters a small limit order is likely to be executed before $T$, but a limit order for $S=1000$ shares is unlikely to be fully filled. All pricing parameters and penalty costs were the same as in the previous example from the section \ref{ssec.fragm}. Substituting these parameters to (\ref{sol.1exchg}) we found that the optimal solution is $(M^{\star},L^{\star})=(728,272)$ shares. Numerical solutions $\hat X^\star$ were then computed for five starting points $X_0$ with a progressively larger number of iterations $N$. For each choice of $X_0$ and $N$ we also estimated an average cost per share $W(\hat X^\star)$ using an additional $L=1000$ samples of $\xi$ generated after $\hat X^\star$ is estimated. Figure \ref{fig.convergence_x} shows that numerical solutions converge to $X^\star$ regardless of the initial point $X_0$ and moreover when $X_0=X^\star$ the iterates remained close to the optimal point. Convergence was also quite fast - after as few as 50 samples the algorithm is within 2\% of the optimal objective value. In the worst case of initial points on the boundary it can take a few thousand samples for the algorithm to converge. It is also worth noting that convergence in terms of the objective value occured significantly faster than convergence in terms of the order allocation vector. Second, we estimated cost savings from dividing orders among multiple trading venues as the number of venues $K$ increases. The cost savings from using our solution were compared with several simple benchmarks. Denote a pure market order allocation by $X_M = (S,0,\dots,0)$, a single limit order allocation by $X_L=(0,S,0,\dots,0)$ and an equal split allocation by $X_E = (\dfrac{S}{K+1},\dfrac{S}{K+1},\dots,\dfrac{S}{K+1})$. Table 2 presents numerical solutions with $X_0=X_E, N=1000$ and average costs per share with additional $L=1000$ test samples for different order sizes $S$ and a different number of exchanges $K=1,\dots,5$. The parameters $s,f,r,\lambda_u,\lambda_o$ were the same as in the section \ref{ssec.fragm}. In this simulation the $K$ exchanges were identical to each other in terms of rebates and initial queue sizes: $r_k=r=20$ mills, $Q_k=Q=2000$ shares, and the order flows were generated according to the single-factor model from section \ref{ssec.fragm} with $\alpha=0.6$. Our results show that the algorithm outperforms the benchmarks, especially when the target quantity $S$ is relatively small. This is because small quantities can be efficiently allocated in form of limit orders among available exchanges and can fully capture limit order cost savings. In contrast, larger quantities are mostly traded with market orders and the difference between $\hat X^\star$ and $X_M$ was small. Comparing the average cost per share $W(X_L)$ and $W(X_E)$ we also see that splitting limit orders across multiple exchanges, even in a naive way, can be very advantageous because limit order fills are not perfectly correlated. Since multiple exchanges in this example have identical initial queues and rebates, the algorithm splits the total limit order quantity equally among them. But the numerical solutions are not the same as the equal-split strategy $X_E$ because the latter also sets $M=\frac{S}{K+1}$, which may be too big or too small depending on problem inputs. Another interesting feature of the numerical solution $\hat X^\star$ is its tendency to oversize the total quantity of orders: $M+\sum_{k=1}^K{L_k}>S$ for $S = 1000, 5000$ and $K = 4,5$. By submitting large orders to multiple exchanges the algorithm reduces the probability of falling behind the target quantity with a relatively low probability of exceeding it. \begin{table}[!h] \begin{center} \small \begin{tabular}{|r|r r r r r r|r r r r|} \hline \multirow{2}{*}{$K$ } & \multicolumn{6}{|c|}{Order allocation} & \multicolumn{4}{|c|}{Average cost, in bp per share}\\ \cline{2-11} & $\big(\hat M^\star$ & $\hat L^\star_1$ & $\hat L^\star_2$ & $\hat L^\star_3$ & $\hat L^\star_4$ & $\hat L^\star_5\big)/S$ & $W(X_M)$ & $W(X_L)$ & $W(X_E)$ & $W(\hat X^\star)$\\ \hline \multicolumn{11}{|c|}{$S=500$}\\ \hline 1 & 0.518 & 0.482 & & & & & 230.00 & 67.87 & 53.84 & 54.05\\ 2 & 0.018 & 0.534 & 0.546 & & & & 230.00 & 66.87 & -64.94 & -124.05\\ 3 & 0.001 & 0.352 & 0.352 & 0.353 & & & 230.00 & 67.98 & -107.16 & -212.63\\ 4 & 0.001 & 0.253 & 0.248 & 0.247 & 0.266 & & 230.00 & 68.06 & -129.95 & -218.97\\ 5 & 0.000 & 0.195 & 0.201 & 0.189 & 0.202 & 0.223 & 230.00 & 68.44 & -144.99 & -219.52\\ \hline \multicolumn{11}{|c|}{$S=1000$}\\ \hline 1 & 0.742 & 0.258 & & & & & 230.00 & 164.06 & 149.06 & 142.16\\ 2 & 0.498 & 0.333 & 0.333 & & & & 230.00 & 163.90 & 57.86 & 53.56\\ 3 & 0.260 & 0.282 & 0.285 & 0.283 & & & 230.00 & 164.34 & -32.93 & -33.57\\ 4 & 0.023 & 0.277 & 0.284 & 0.278 & 0.287 & & 230.00 & 163.55 & -104.87 & -123.95\\ 5 & 0.001 & 0.228 & 0.227 & 0.226 & 0.225 & 0.229 & 230.00 & 163.81 & -138.32 & -191.74\\ \hline \multicolumn{11}{|c|}{$S=5000$}\\ \hline 1 & 0.902 & 0.098 & & & & & 230.00 & 240.82 & 225.82 & 213.75\\ 2 & 0.832 & 0.135 & 0.135 & & & & 230.00 & 240.86 & 211.70 & 196.74\\ 3 & 0.779 & 0.159 & 0.159 & 0.159 & & & 230.00 & 240.77 & 194.81 & 178.94\\ 4 & 0.733 & 0.183 & 0.183 & 0.183 & 0.183 & & 230.00 & 240.84 & 177.30 & 161.31\\ 5 & 0.688 & 0.167 & 0.167 & 0.167 & 0.167 & 0.167 & 230.00 & 240.86 & 159.33 & 143.68\\ \hline \end{tabular} \normalsize \vskip 4pt \caption{Optimal order allocations for different number of exchanges and average costs for different order allocation strategies.} \end{center} \end{table} To further illustrate the structure of a numerical solution we performed a sensitivity analysis. The base case for this analysis was the example with $K=2$ exchanges from the section \ref{ssec.fragm} with a correlation parameter set to $\alpha=0.6$ . Varying one of the parameters at a time, such as an initial queue size or a rebate on one of the exchanges, we plotted the numerical solution $\hat X^\star$ after $N=1000$ iterations, together with an analytical solution for a single exchange. The results are presented on Figures \ref{fig.twoex_grp1} and \ref{fig.twoex_grp2}. Similarly to solutions for a single exchange, limit order sizes on two exchanges $L_1,L_2$ decrease and a market order size $M$ increases as the penalty $\lambda_u$ increases. Increasing the half-spread $h$, the rebate $r_1$ or the fee $f$ makes a limit order on exchange number one more attractive, so $L_1$ increases and $M$ decreases. Because the penalty $\lambda_u$ is large in this example relative to fee and rebate values, changes in the queue size $Q_1$ and the order outflow mean $\mu_1$ have a much stronger effect on the optimal solution than $r_1$ and $f$. Both decreasing the $Q_1$ and increasing $\mu_1$ make a limit order fill more likely at exchange number one and $L_1$ increases. Finally, as in the case of a single exchange, the target size $S$ has a strong effect on the optimal order allocation. While $S$ remains small, only limit orders are used, but as it becomes larger it is difficult to fill that quantity solely with limit orders and the optimal market order size begins to grow to limit the execution risk. \begin{figure}[h!] \noindent \begin{center} { \includegraphics[width=120mm]{convergence_x} } \end{center} \caption{Convergence of numerical solutions and objective values to an optimal point from different initial points.} \label{fig.convergence_x} \end{figure} \section{Conclusion}\label{sec.conc} We have formulated the {\it optimal order placement problem} for a market participant able to submit market orders and limit orders across multiple exchanges, and studied its solution properties using historical data and numerical simulations. In the case when only one exchange is available we have derived an optimal split between limit and market orders and showed that an optimal order allocation depends on trader's aversion to execution risk. For the general case of multiple exchanges, we provide a characterization of the optimal order placement strategy in terms of execution shortfall probabilities. To solve the problem in practical applications we propose a fast and straightforward numerical algorithm that re-samples past order fill data to optimize future order executions. Using this algorithm, we have studied the sensitivities of an optimal order allocation to problem inputs and showed that a simultaneous placement of limit orders on multiple trading venues according to our method can lead to a substantial reduction of transaction costs. \newpage \begin{landscape} \thispagestyle{empty} \begin{figure}[h!] \noindent \begin{center} { \includegraphics[width=200mm]{twoex_grp1} } \end{center} \caption{Sensitivity analysis for a numerical solution $\hat X^{\star}=(M,L_1,L_2)$ with two exchanges and an optimal solution $(M^a,L^a)$ with the first exchange only.} \label{fig.twoex_grp1} \end{figure} \end{landscape} \newpage \begin{landscape} \thispagestyle{empty} \begin{figure}[h!] \noindent \begin{center} { \includegraphics[width=200mm]{twoex_grp2} } \end{center} \caption{Sensitivity analysis for a numerical solution $\hat X^{\star}=(M,L_1,L_2)$ with two exchanges and an optimal solution $(M^a,L^a)$ with the first exchange only.} \label{fig.twoex_grp2} \end{figure} \end{landscape} \begin{APPENDIX}{Proofs} \noindent{\bf Proposition 1} {\it Consider $\mathcal{C}$ - a compact convex subset of $\mathbb{R}^{K+1}_+$ defined by $$\mathcal{C}\overset{\Delta}{=}\left\{X\in\mathbb{R}^{K+1}_+\Bigm| 0\leq M\leq S,~~0\leq L_k\leq S-M, k=1,\dots,K, ~~M+\sum_{k=1}^K{L_k}\geq S\right\}$$ \noindent Under assumptions A1-A2 for any $\tilde X\notin\mathcal{C}$, $\exists \tilde X'\in\mathcal{C}$ with $V(\tilde X')\leq V(\tilde X)$. Moreover, if $\underset{k}{\min}\left\{\mathbb{P}(\xi_k>Q_k+S)\right\}>0$, the inequality is strict: $V(\tilde X')<V(\tilde X)$. } \newline \noindent\textbf{Proof:} First, for any allocation $\tilde X$ that has $\tilde M>S$, we automatically have $A(\tilde X)>S$ and we can show that the (random) cost and penalty of $\tilde X$ are larger than those of $X^{naive}\overset{\Delta}{=}(S,0,\dots,0)\in\mathcal{C}$: \begin{multline} v({\tilde X},\xi) - v(X^{naive},\xi)= (h+f)(\tilde M-S)-\sum_{k=1}^K(h+r_k)((\xi_k-Q_k)_+-(\xi_k-Q_k-L_k)_+)+\\ \lambda_o\left(\tilde M-S+\sum_{k=1}^K\left((\xi_k-Q_k)_+-(\xi_k-Q_k-L_k)_+\right)\right)= \\ (\lambda_o+h+f)(\tilde M-S)+\sum_{k=1}^K(\lambda_o-h-r_k)((\xi_k-Q_k)_+-(\xi_k-Q_k-L_k)_+)>0, \notag \end{multline} \noindent which holds for all random $\xi$. Therefore, $V(\tilde X)>V(X^{naive})$. Similarly, for any allocation $\tilde X$ with $\tilde L_k>S - \tilde M$ define a different allocation $\tilde X'$ by $\tilde M'=\tilde M$, $\tilde L'_j=\tilde L_j, \forall j\neq k$ and $\tilde L'_k=S - \tilde M$. Then $v(\tilde X,\xi) - v(\tilde X',\xi)=0$ on the event $B~=~\left\{\omega|\xi_k(\omega)<Q_k+S - M\right\}$. On its complementary event $B^c$, $$v(\tilde X,\xi)-v(\tilde X',\xi)=-(h+r_k)((\xi_k-Q_k-S+\tilde M)_+-(\xi_k-Q_k-\tilde L_k)_+)$$ $$+\lambda_o((\xi_k-Q_k-S+\tilde M)_+-(\xi_k-Q_k-\tilde L_k)_+).$$ Therefore \begin{multline} V(\tilde X) - V(\tilde X') = \mathbb{E}\left[v(\tilde X,\xi)- v(\tilde X',\xi)|B\right]\mathbb{P}(B)+ \mathbb{E}\left[v(\tilde X,\xi)-v(\tilde X',\xi)|B^c\right]\mathbb{P}(B^c) = \\ 0+\mathbb{E}\left[(\lambda_o-(h+r_k))((\xi_k-Q_k-S+\tilde M)_+-(\xi_k-Q_k-\tilde L_k)_+)|B^c\right]\mathbb{P}(B^c)\geq 0 \notag \end{multline} \noindent with a strict inequality if $\mathbb{P}(B^c)>0$. If $\tilde X'\notin\mathcal{C}$, we can continue truncating limit order sizes $\tilde L'_j>S - \tilde M'$ following the same argument. Each time the truncation does not increase the objective function and finally we obtain $\tilde X''\in\mathcal{C}$, such that $V(\tilde X'')\leq V(\tilde X)$. Next, if $\tilde X$ is such that $\tilde M-\sum_{k=1}^K{\tilde L_k}<S$ define $s=S-\tilde M-\sum_{k=1}^K{\tilde L_k}$, take $\tilde M'=\tilde{M}, \tilde L'_k=\tilde{L}_k, k=1,\dots,K-1$ and $\tilde L'_K=\tilde{L}_k+s$. Then, on the event $B=\left\{\omega|\xi_K(\omega)<Q_K+\tilde L_K\right\}$ we have $v(\tilde X,\xi)= v(\tilde X',\xi)$. However, on the event $B^c$, $$v(\tilde X,\xi)-v(\tilde X',\xi)=(h+r_K)((\xi_K-Q_K-\tilde{L}_K)_+-(\xi_k-Q_k-\tilde{L}_K-s)_+)+\lambda_u((\xi_K-Q_K-\tilde{L}_K)_+-(\xi_k-Q_k-\tilde{L}_K-s)_+),$$ therefore \begin{multline} V(\tilde X) - V(\tilde X') = \mathbb{E}\left[v(\tilde X,\xi) - v(\tilde X',\xi)|B\right]\mathbb{P}(B)+ \mathbb{E}\left[v(\tilde X,\xi)- v(\tilde X',\xi) |B^c\right]\mathbb{P}(B^c) = \\ 0+\mathbb{E}\left[(\lambda_u+(h+r_k))((\xi_K-Q_K-\tilde{L}_K)_+-(\xi_k-Q_k-\tilde{L}_K-s)_+)|B^c\right] \mathbb{P}(B^c)\geq 0 \notag \end{multline} \noindent with a strict inequality if $\mathbb{P}(B^c)>0$. $\blacksquare$ \newline \noindent{\bf Proposition 2} {\it Under assumptions A1-A2, $V(X)$ is a convex function on $\mathbb{R}^{K+1}_+$, bounded from below and admits a global minimizer $X^\star\in\mathcal{C}$. }\newline \noindent \textbf{Proof:} First, note that $(\xi_k-Q_k)_+-(\xi_k-Q_k-L_k)_+$ are concave functions of $L_k$. Therefore, $A(X,\xi)$ is concave as a sum of concave functions. Similarly, the cost term in $v(X,\xi)$ is a sum of convex functions, as long as $r_k\geq -h, k=1,\dots,K$ and is itself a convex function. Second, since $S - A(X,\xi)$ is a convex functon of $X$, and the function $h(x)\overset{\Delta}{=}\lambda_u(x)_+-\lambda_o(-x)_+$ is convex in $x$ for positive $\lambda_u,\lambda_o$, so the penalty term $h\left(S - A(X,\xi))\right)$ is also convex. If $\lambda_o > h+\underset{k}{\max}\{r_k\}$ the function $V(X)$ is also bounded from below since $v(X,\xi)\geq-(h+\underset{k}{\max}\{r_k\})S$. Finally, since $V(X)$ is convex, it is also continous and reaches a local minimum $V_{min}$ on the compact set $\mathcal{C}$ at some point $X^\star\in\mathcal{C}$. By convexity, $V_{min}$ is a global minimum of $V(X)$ on $\mathcal{C}$. Moreover, since $\lambda_o > h+\underset{k}{\max}\{r_k\}$, Proposition \ref{prop.lubound} guarantees that $V_{min}<V(\tilde X)$ for any $\tilde X\notin\mathcal{C}$, so $V_{min}$ is also a global minimum of $V(X)$ on $\mathbb{R}^{K+1}_+$. $\blacksquare$ \newline \noindent{\bf Proposition 3} {\it Assume that $\xi$ has a continuous distribution and (A1-A2) hold. The optimal order allocation depends on $\lambda_u$: If $\lambda_u\leq\underline{\lambda_u}\overset{\Delta}{=}\dfrac{2h+f+r}{F(Q+ S)}-(h+r)$, the optimal allocation is $(M^{\star},L^{\star})=(0,S)$. If $\lambda_u\geq\overline{\lambda_u}\overset{\Delta}{=}\dfrac{2h+f+r}{F(Q)}-(h+r)$, the optimal allocation is $(M^{\star},L^{\star})=(S,0)$. If $\lambda_u\in(\underline{\lambda_u},\overline{\lambda_u})$, the optimal allocation is given by (\ref{sol.1exchg}). } \noindent {\bf Proof:} By Proposition \ref{prop.lubound} there exists an optimal split $ (M^\star,L^\star)\in\mathcal{C}$ between limit and market orders. Moreover for $K=1$ the set $\mathcal{C}$ reduces to a line $M^\star+L^\star=S$ so it is sufficient to find $M^\star$. Restricting $L=S-M$ implies that $\{A(X,\xi)>S\}=\varnothing$, $\{A(X,\xi)<S, \xi>Q+L\}=\varnothing$, and we can rewrite the objective function as \begin{multline}\label{obj.onex} V(M)=\mathbb{E}\Bigl[(h+f)M-(h+r)((\xi-Q)_+-(\xi-Q-S+M)_+)~~+\\ \lambda_u\left(S - M-((\xi-Q)_+-(\xi-Q-S+M)_+)))\right)_+\Bigr]. \end{multline} \noindent For $M\in(0,S)$ the expression under the expectation in (\ref{obj.onex}) is bounded for all $\xi$ and differentiable with respect to $M$ for almost all $\xi$, so we can compute $V'(M)=\frac{dV(M)}{dM}$ by interchanging the order of differentiation and integration (see e.g. \citet{Aliprantis1998}, Theorem 24.5): \begin{equation} V'(M)=\mathbb{E}\Bigl[h+f+(h+r)\indicator{\xi>Q+S-M}-\lambda_u\indicator{\xi<Q+S-M}\Bigr]=2h+f+r-(h+r+\lambda_u)F(Q+S-M) \end{equation} \noindent Note that if $\lambda_u\leq\dfrac{2h+f+r}{F(Q+S)}-(h+r)$, then $V'(M)\geq 0$ for $M\in(0,S)$ and therefore $V$ is non-decreasing at these points. Checking that $V(S)-V(0)\geq(h+f-\lambda_u)S+(\lambda_u+h+r)S(1-F(Q+S))\geq 0$ we conclude that $M^\star = 0$. Similarly, if $\lambda_u\geq\dfrac{2h+f+r}{F(Q)}-(h+r)$, then $v(M)\leq 0$ for all $M\in(0,S)$ and $V(M)$ is non-increasing at these points. Checking that $V(S)-V(0)\leq(h+f-\lambda_u)S+(\lambda_u+h+r)S(1-F(Q))\leq 0$ we conclude that $M^\star = S$. Finally, if $\lambda_u$ is between these two values, $\exists\epsilon>0$, such that $V'(\epsilon)<0, V'(S-\epsilon)>0$ and by continuity of $V'$ there is a point where $V'(M^\star)=0$. This $M^\star$ is optimal by convexity of $V(M)$ and (\ref{sol.1exchg}) solves equations $v(M^\star,\xi)=0, L^\star=S-M^\star$. $\blacksquare$ \newline \noindent{\bf Proposition 4} {\it Assume (A1-A2), also assume that the distribution of $\xi$ is continuous, $\underset{k}{\max}\left\{F_k(Q_k+S)\right\}<1$ and $\lambda_u<\underset{k}{\max}\left\{\dfrac{2h+f+r_k}{F_k(Q_k)}-(h+r_k)\right\}$. Then: \vskip 6pt \begin{enumerate} \item Any optimal allocation $X^\star$ has a positive market order quantity $M^\star>0$ if \eqref{cond.mrkord} holds. \item Any optimal allocation $X^\star$ has a positive limit order quantity $L^\star_j>0$ if \eqref{cond.limord} holds. \item If \eqref{cond.mrkord}-\eqref{cond.limord} hold for all exchanges $j=1,\dots,K$, a necessary and sufficient condition for optimality of an order allocation $X^\star\in\mathcal{C}$ is that it solves equations \eqref{FOC1}--\eqref{FOC2}. \end{enumerate} } \noindent {\bf Proof:} Proposition \ref{prop.objfun} implies the existence of an optimal order allocation $X^\star\in\mathcal{C}$. First, we define $X_M\overset{\Delta}{=}(S,0,\dots,0)$ and prove that $X^\star\neq X_M$ by contradiction. If $X_M$ were optimal in the problem (\ref{optim.set}) it would also be optimal in the same problem with a constraint $L_k=0,k\neq j$, for any one $j$. In other words, the solution $(S, 0)$ would be optimal for any one-exchange problem, defined by using only exchange $j$. But by our assumption, there exists$ J$ such that $\lambda_u<\dfrac{2h+f+r_J}{F_J(Q_J)}-(h+r_J)$ and Proposition \ref{prop.1exchg} implies that $(S, 0)$ is not optimal for the $J$-th single-exchange subproblem, leading to a contradiction. The function $v(X,\xi)$ is bounded for $X\in\mathcal{C}$ and for all $\xi$, differentiable with respect to $M$ and $L_k, k=1,\dots,K$ for $X\in\mathcal{C}\backslash\left\{X_M\right\}$ for almost all $\xi$. Applying the same theorem as in the proof of Proposition \ref{prop.1exchg} we conclude that $V(X)$ is differentiable for $X\in\mathcal{C}\backslash\left\{X_M\right\}$ and we can compute all of its partial derivatives by interchanging the order of differentiation and integration. The KKT conditions for problem (\ref{optim.set}) and $X\in\mathcal{C}\backslash\left\{X_M\right\}$ are \begin{align} h + f -\lambda_u\mathbb{P}(A(X^\star,\xi)<S)& + \lambda_o\mathbb{P}(A(X^\star,\xi)>S)-\mu_0=0 \label{KKT1}\\ -(h+r_k)\mathbb{P}(\xi_k>Q_k+L^{\star}_k)& -\lambda_u\mathbb{P}(A(X^\star,\xi)<S, \xi_k>Q_k+L^{\star}_k) + \notag\\ & \lambda_o\mathbb{P}(A(X^\star,\xi)>S, \xi_k>Q_k+L^{\star}_k)-\mu_k=0,~~k=1,\dots, K \label{KKT2}\\ M\geq 0,~~L_k\geq 0,~~\mu_0\geq 0,&~~\mu_k\geq 0,~~ \mu_0M=0,~~\mu_kL_k=0,~~k=1,\dots, K \label{KKT3} \end{align} Since the objective function $V(\cdot)$ is convex, conditions \eqref{KKT1}--\eqref{KKT3} are both necessary and sufficient for optimality. The first result of this proposition follows from considering any $\tilde X$ with $\tilde M=0$: \noindent $V(\tilde X)\geq\lambda_u S\mathbb{P}\left(\underset{k}{\bigcap}\{\xi_k\leq Q_k\}\right)-(h+\underset{k}{\max}\{r_k\})S\mathbb{P}\left(\overline{\underset{k}{\bigcap}\{\xi_k\leq Q_k\}}\right)\geq(h+f)S=V(X_M)$ and we already argued that $\exists X^\star$ with $V(X^\star)<V(X_M)$, so $X^\star\neq\tilde X$ and therefore $M^\star>0$. Rearranging terms in a $j$-th equality (\ref{KKT2}) we obtain \begin{equation}\label{KKT2supp} \mathbb{P}(\xi_j>Q_j+L^{\star}_j)\left[\lambda_o-(h+r_j) -(\lambda_u+\lambda_o)\mathbb{P}(A(X^\star,\xi)<S | \xi_j>Q_j+L^{\star}_j)\right]-\mu_j=0 \end{equation} \noindent The term in square brackets in (\ref{KKT2supp}) is negative for any $X\in\mathcal{C}\backslash\left\{X_M\right\}$ with $L_j=0$, because \noindent $\mathbb{P}(A(X,\xi)<S | \xi_j>Q_j+L_j)> \mathbb{P}\left(\underset{k\neq j}{\bigcap}\{\xi_k\leq Q_k\}\bigg|\xi_j>Q_j\right)>\frac{\lambda_o-(h+r_j)}{\lambda_u+\lambda_o}$ \noindent by assumption and since $\mu_j\geq 0$ the condition (\ref{KKT2}) cannot be satisfied with $L^{\star}_j=0$. We showed that $M^{\star}>0, L^{\star}_j>0$ for all $j=1,\dots,K$ and therefore, $\mu_0=\mu_1=\dots=\mu_K=0$ by complimentary slackness. Then the KKT conditions \eqref{KKT1}--\eqref{KKT3} reduce to \eqref{FOC1}--\eqref{FOC2}. $\blacksquare$ \noindent{\bf Corollary} {\it Consider the case of two exchanges with $\xi_1,\xi_2$ that are independent and have continuous distributions. If \begin{enumerate} \item $\underset{k=1,2}{\max}\left\{F_k(Q_k+S)\right\}<1$ \item $\lambda_u<\underset{k=1,2}{\max}\left\{\dfrac{2h+f+r_k}{F_k(Q_k)}-(h+r_k)\right\}$, $\lambda_u\geq\dfrac{2s+f+\underset{k=1,2}{\max}\{r_k\}}{F_1(Q_1)F_2(Q_2)}-(h+\underset{k=1,2}{\max}\{r_k\})$ \item $F_1(Q_1)<1-\dfrac{h+r_2}{\lambda_o}, F_2(Q_2)<1-\dfrac{h+r_1}{\lambda_o}$ \end{enumerate} then there exists an optimal order allocation $X^\star=(M^{\star}, L^{\star}_1, L^{\star}_2)\in int\{\mathcal{C}\}$ and it verifies (\ref{sol.2exchg.1}-\ref{sol.2exchg.m}). }\newline \noindent {\bf Proof:} Solutions on the boundary of $\mathcal{C}$ are sub-optimal: $M^\star = 0$ and $M^\star = S$ are ruled out by assumption 2, $L^\star_1 = S-M$ and $L^\star_2 = S-M$ are ruled out by assumption 3 and (\ref{KKT2}). Solutions with $M^\star+\sum\limits_{k=1}^KL^\star_k = S$ are ruled out by directly checking (\ref{KKT2}). Finally, $L^\star_1 = 0$ and $L^\star_2 = 0$ are also ruled out by (\ref{KKT2}). For example if $L^\star_1 = 0$, then by Proposition \ref{prop.lubound} $M^\star+L^\star_2 = S$ and in (\ref{KKT2}) $\mu_2=0$ by complimentary slackness, $\mathbb{P}(A(X^\star,\xi)<S, \xi_2>Q_2+L^{\star}_2)=\mathbb{P}(A(X^\star,\xi)>S, \xi_2>Q_2+L^{\star}_2)=0$. But then (\ref{KKT2}) cannot hold because $\mathbb{P}(\xi_2>Q_2+L^{\star}_2)>0$. For any $X\in int\{\mathcal{C}\}$, $A(X,\xi)>S$ if and only if all the following three inequalities are satisfied: \begin{subequations} \begin{align} & \xi_1 > Q_1 + S - M - L_2 \label{overfill.2exchg.1}\\ & \xi_2 > Q_2 + S - M - L_1 \label{overfill.2exchg.2}\\ & \xi_1 + \xi_2 > Q_1 + Q_2 + S - M \label{overfill.2exchg.3} \end{align} \end{subequations} \noindent These inequalities give a simple characterization of the event $\{A(X,\xi)>S\}$ which is directly verified by considering subsets of $(\xi_1,\xi_2)$ forming a complete partition of $\mathbb{R}^2_+$. {\bf Case 1:} $\xi_1>Q_1+L_1, \xi_2>Q_2+L_2$. Since $L_1+L_2+M>S$, we have $A(X,\xi)=L_1+L_2+M>S$ and at the same time all of the inequalities (\ref{overfill.2exchg.1}-\ref{overfill.2exchg.3}) are satisfied, so they are trivially equivalent in this case. {\bf Case 2:} $\xi_1>Q_1+L_1, Q_2\leq\xi_2\leq Q_2+L_2$. Because of the condition $\xi_1>Q_1+L_1$, (\ref{overfill.2exchg.1}) is satisfied. We have in this case that $A(X,\xi)=L_1+\xi_2-Q_2+M$ and thus $A(X,\xi)>S$ if and only if (\ref{overfill.2exchg.2}) is satisfied. Finally, $\xi_1>Q_1+L_1$ together with (\ref{overfill.2exchg.2}) imply (\ref{overfill.2exchg.3}), so $A(X,\xi)>S$ and (\ref{overfill.2exchg.1}-\ref{overfill.2exchg.3}) are equivalent in this case. {\bf Case 3:} $\xi_2>Q_2+L_2, Q_1\leq\xi_1\leq Q_1+L_1$. Similarly to Case 2 we can show that inequalities (\ref{overfill.2exchg.1}-\ref{overfill.2exchg.3}) are satisfied if and only if $A(X,\xi)>S$. {\bf Case 4:} $Q_1+S-M-L_2<\xi_1\leq Q_1+L_1, Q_2+S-M-L_1<\xi_2\leq Q_2+L_2$. This set is non-empty because $0<S - M- L_1 <L_2$ and similarly for $L_1, L_2$ reversed. Inequalities \eqref{overfill.2exchg.1}--\eqref{overfill.2exchg.2} hold trivially, only (\ref{overfill.2exchg.3}) needs to be checked. We can write $A(X,\xi)=\xi_1-Q_1+\xi_2-Q_2+M>S$ if and only if (\ref{overfill.2exchg.3}) holds, so $A(X,\xi)>S$ is equivalent to (\ref{overfill.2exchg.1}-\ref{overfill.2exchg.3}). {\bf Case 5:} Outside of Cases 1-4, either (\ref{overfill.2exchg.1}) or (\ref{overfill.2exchg.2}) is not satisfied. If $\xi_1\leq Q_1+S-M-L_2, \xi_2\leq Q_2+L_2$, then $A(X,\xi)\leq S-M-L_2 + L_2 +M = S$. The case $\xi_2\leq Q_2+S-M-L_1, \xi_1\leq Q_1+L_1$ is completely symmetric, and it shows that neither $A(X,\xi)> S$ nor (\ref{overfill.2exchg.1}-\ref{overfill.2exchg.3}) hold in this case. \noindent Next, we use inequalities (\ref{overfill.2exchg.1}-\ref{overfill.2exchg.3}) to characterize the set $\{A(X,\xi)>S\}$ in the first-order conditions \eqref{FOC1}--\eqref{FOC2}. We observe that in the two-exchange case \begin{align*} & \{A(X,\xi)>S,\xi_1>Q_1+L_1\}=\{\xi_1>Q_1+L_1,\xi_2>Q_2+S-M-L_1\}\\ & \{A(X,\xi)>S,\xi_2>Q_2+L_2\}=\{\xi_2>Q_2+L_2,\xi_1>Q_1+S-M-L_2\}, \end{align*} \noindent and then use the independence of $\xi_1$ and $\xi_2$ to compute \begin{align*} & \mathbb{P}(A(X,\xi)>S|\xi_1>Q_1+L_1)=\bar F_2(Q_2+S-M-L_1) \\ &\mathbb{P}(A(X,\xi)>S|\xi_2>Q_2+L_2)=\bar F_1(Q_1+S-M-L_2) \end{align*} \noindent Together with \eqref{FOC1} and \eqref{FOC2}, this leads to a pair of equations for limit orders sizes: \begin{align*} \bar F_2(Q_2+S-M-L_1)=\frac{\lambda_u+h+r_1}{\lambda_u+\lambda_o} \quad \bar F_1(Q_1+S-M-L_2)=\frac{\lambda_u+h+r_2}{\lambda_u+\lambda_o} \end{align*} \noindent whose solution is given by $L^{\star}_1, L^{\star}_2$ from (\ref{sol.2exchg.1},\ref{sol.2exchg.2}). To obtain the equation (\ref{sol.2exchg.m}), we rewrite the first equation in (\ref{FOC1},\ref{FOC2}) using the inequalities (\ref{overfill.2exchg.1}-\ref{overfill.2exchg.3}). Then ${P}(A(X,\xi)>S)$ may be computed as the integral of the product measure $F_1 \otimes F_2 $ over the region defined by $$ U(Q,S,M,L_1,L_2)= \{ (x_1,x_2)\in \mathbb{R}^2,\quad x_1>Q_1+ S - M-L_2,\quad x_2>Q_2+ S - M-L_1,\quad x_1+x_2>Q_1+Q_2+S-M \}.$$ This integral is given by \begin{align*} &{P}(A(X,\xi)>S)=F_1 \otimes F_2 \left( U(Q,S,M,L_1,L_2)\right)\\ &=\bar F_1( Q_1+L_1)\bar F_2(Q_2+S-M-L_1) +\int\limits_{Q_1+ S-M-L_2}^{Q_1+L_1}{\bar F_2(Q_1+Q_2+S-M-x_1)dF_1(x_1)}=\frac{\lambda_u-(h+f)}{\lambda_u+\lambda_o} \end{align*} $\blacksquare$ \end{APPENDIX} \newpage \bibliographystyle{ormsv080}
\section{Introduction} The stochastic heat equation \begin{align}\label{E2:Heat} \begin{cases} \displaystyle \left(\frac{\partial }{\partial t} - \frac{\nu}{2} \frac{\partial^2 }{\partial x^2}\right) u(t,x) = \rho(u(t,x)) \:\dot{W}(t,x),& x\in \mathbb{R},\; t \in\mathbb{R}_+^*, \\ \displaystyle \quad u(0,\cdot) = \mu(\cdot)\;, \end{cases} \end{align} \nomenclature{$\dot{W}$}{Space-time white noise \nomrefpage}\noindent where $\dot{W}$ is space-time white noise, $\rho(u)$ is globally Lipschitz, $\mu$ is the initial data, and $\mathbb{R}_+^*=\;]0,+\infty[$, has been intensively studied during last two decades by many authors \cite{BertiniCancrini94Intermittence,CarmonaMolchanov94,ConusEct12Initial, ConusKhosh10Farthest, ConusKhosh10Weak,DalangKhNualart07HittingAdditive,DalangKhNualart09HittingMult, FoondunKhoshnevisan08Intermittence,Mueller91Support,PospisilTribe07Parameter, SanSoleSarra99Holder, SanSoleSarra99Path,Shiga94Two,Walsh86} . In particular, the special case $\rho(u)=\lambda u$ is called {\it the parabolic Anderson model} \cite{CarmonaMolchanov94}. Our work focuses on this equation with general deterministic initial data, and we study how the initial data affects the solution. The one-dimensional heat kernel function is \begin{align}\label{E2:G-Heat} G_\nu(t,x) := \frac{1}{\sqrt{2\pi \nu t}} \exp\left\{-\frac{|x|^2}{2\nu t}\right\},\quad (t,x)\in\mathbb{R}_+^*\times\mathbb{R}\:. \end{align} \nomenclature{$G_\nu(t,x)$}{heat kernel function \nomrefeqpage}\noindent For the existence of random field solutions to \eqref{E2:Heat}, the case where the initial data $\mu$ is a bounded and measurable function is covered by the classical theory of Walsh \cite{Walsh86}. When $\mu$ is a positive Borel measure on $\mathbb{R}$ such that \begin{align}\label{E2:BC-InitD} \sup_{t\in [0,T]} \sup_{x\in\mathbb{R}} \sqrt{t} \left(\mu*G_\nu(t,\circ)\right)(x)<\infty,\quad\text{for all $T>0$}, \end{align} where $*$ denotes convolution in the spatial variable, Bertini and Cancrini \cite{BertiniCancrini94Intermittence} gave an ad-hoc definition for the Anderson model via a smoothing of the space-time white noise and a Feynman-Kac type formula. Their analysis depends heavily on properties of the local times of Brownian bridges. Recently, Conus and Khoshnevisan \cite{ConusKhosh10Weak} constructed a weak solution defined through certain norms on random fields. The initial data has to verify certain technical conditions, which include the Dirac delta function in some of their cases. In particular, the solution is defined for almost all $(t,x)$, but not at specific $(t,x)$. More recently, Conus, Joseph, Khoshnevisan and Shiu \cite{ConusEct12Initial} also studied random field solutions. In particular, they require the initial data to be a finite measure of compact support. We improve the existence result by working under a much weaker condition on initial data, namely, $\mu$ can be any signed Borel measure over $\mathbb{R}$ such that \begin{align}\label{E2:J0finite} \left(|\mu| * G_\nu(t,\cdot)\right) (x)<+\infty\;, \quad \text{for all $t>0$ and $x\in\mathbb{R}$}\;, \end{align} \nomenclature{$*$}{space convolution \nomrefeqpage}\noindent where, from the Jordan decomposition, $\mu=\mu_+-\mu_-$ where $\mu_\pm$ are two non-negative Borel measures and $|\mu|:= \mu_++\mu_-$. For instance, if $\mu(\ensuremath{\mathrm{d}} x)=f(x)\ensuremath{\mathrm{d}} x$, then $f(x)=\exp\left(a|x|^p\right)$, $a>0$, $p\in \;]0,2[$, (i.e., exponential growth at $\pm \infty$), will satisfy this condition. Proposition \ref{P2:D-Delta} below shows that initial data cannot be extended beyond measures to other Schwartz distributions, even with compact support. Moreover, we obtain estimates for the moments $\mathbb{E}(|u(t,x)|^p)$ with both $t$ and $x$ fixed for all even integers $p\ge 2$. In particular, for the parabolic Anderson model, we give an explicit formula for the second moment of the solution. When the initial data is either Lebesgue measure or the Dirac delta function, we give explicit formulae for the two-point correlation functions (see \eqref{E2:TP-Lebesgue} and \eqref{E2:TP-Delta} below), which can be compared to the integral form in Bertini and Cancrini's paper \cite[Corollaries 2.4 and 2.5]{BertiniCancrini94Intermittence}. Our proof of existence is based on the standard Picard iteration. The main difference from the conventional situation is that instead of applying Gronwall's lemma to bound the second moment from above, we show that the sequence of the second moments in the Picard iteration converges to an explicit formula (in the case of the parabolic Anderson model). After establishing the existence of random field solutions, we study whether the solution exhibits intermittency properties. More precisely, define the {\it upper and lower Lyapunov exponents} for constant initial data (Lebesgue measure) as follows \begin{align}\label{E2:Lyapunov} \overline{\lambda}_p:=& \mathop{\lim\sup}_{t\rightarrow+\infty}\frac{\log \mathbb{E}\left[|u(t,x)|^p\right]}{t},\qquad \underline{\lambda}_p:=\mathop{\lim\inf}_{t\rightarrow+\infty}\frac{\log \mathbb{E}\left[|u(t,x)|^p\right]}{t} \;. \end{align} \nomenclature{$\overline{\lambda}_p$}{upper Lyapunov exponent of order $p$ \nomrefeqpage}\noindent \nomenclature{$\underline{\lambda}_p$}{lower Lyapunov exponent of order $p$ \nomrefeqpage}\noindent Following Bertini and Cancrini \cite{BertiniCancrini94Intermittence}, we say that the solution is {\it intermittent} if $\lambda_n:=\underline{\lambda}_n =\overline{\lambda}_n$ and the strict inequalities \begin{align}\label{E2:Intermit} \lambda_1 < \frac{\lambda_2}{2}<\cdots<\frac{\lambda_n}{n}<\cdots \end{align} are satisfied. Carmona and Molchanov gave the following definition \cite[Definition III.1.1, on p. 55]{CarmonaMolchanov94}: \begin{definition}[Intermittency]\label{D2:Intermit} Let $p$ be the smallest integer for which $\lambda_p>0$. When $p<\infty$, we say that the solution $u(t,x)$ shows {\em (asymptotic) intermittency of order $p$} and {\em full intermittency} when $p=2$. \end{definition} They showed that full intermittency implies the intermittency defined by \eqref{E2:Intermit} (see \cite[III.1.2, on p. 55]{CarmonaMolchanov94}). This mathematical definition of intermittency is related to the property that the solutions develop high peaks on some small ``islands". The parabolic Anderson model has been well studied: see \cite{CarmonaMolchanov94} for a discrete approximation and \cite{BertiniCancrini94Intermittence} and \cite{FoondunKhoshnevisan08Intermittence} for the continuous version. Further discussion can be found in \cite{Zeldovich90Almighty}. When the initial data are not homogeneous, in particular, when they have certain exponential decrease at infinity, Conus and Khoshnevisan \cite{ConusKhosh10Farthest} defined the following {\it lower and upper exponential growth indices}: \begin{align} \label{E2:GrowInd-0} \underline{\lambda}(n):= & \sup\left\{\alpha>0: \underset{t\rightarrow \infty}{\lim\sup} \frac{1}{t}\sup_{|x|\ge \alpha t} \log \mathbb{E}\left(|u(t,x)|^n\right) >0 \right\}\;,\\ \label{E2:GrowInd-1} \overline{\lambda}(n) := & \inf\left\{\alpha>0: \underset{t\rightarrow \infty}{\lim\sup} \frac{1}{t}\sup_{|x|\ge \alpha t} \log \mathbb{E}\left(|u(t,x)|^n\right) <0 \right\}\:, \end{align} \nomenclature{$\underline{\lambda}(n)$}{lower exponential growth indices of order $n$ \nomrefeqpage}\noindent \nomenclature{$\overline{\lambda}(n)$}{upper exponential growth indices of order $n$ \nomrefeqpage}\noindent and proved that if the initial data $\mu$ is a non-negative, lower semicontinuous function with compact support of positive measure, then for the Anderson model ($\rho(u)=\lambda u$), \[ \frac{\lambda^2}{2\pi} \le \underline{\lambda}(2) \le \overline{\lambda}(2) \le \frac{\lambda^2}{2}\;. \] We improve this result by showing that $\underline{\lambda}(2)= \overline{\lambda}(2) = \lambda^2/2$, and extend this to more general measure-valued initial data. This is possible mainly thanks to our explicit formula for the second moment. We now discuss the regularity of the random field solution. Denote by $C_{\beta_1,\beta_2}(D)$ the set of random fields whose trajectories are almost surely $\beta_1$-H\"older continuous in time and $\beta_2$-H\"older continuous in space on the domain $D\subseteq \mathbb{R}_+\times\mathbb{R}$, and let \[ C_{\beta_1-,\beta_2-}(D) := \bigcup_{\alpha_1\in \;\left]0,\beta_1\right[} \bigcup_{\alpha_2\in \;\left]0,\beta_2\right[} C_{\alpha_1,\alpha_2}(D)\;. \] \nomenclature{$C_{\beta_1,\beta_2}(D)$}{ the set of random fields whose trajectories are almost surely $\beta_1$-H\"older continuous in time and $\beta_2$-H\"older continuous in space over domain $(t,x)\in D\subseteq \mathbb{R}_+\times\mathbb{R}$ \nomrefpage}\noindent In Walsh's notes \cite[Corollary 3.4, p. 318]{Walsh86}, a slightly different equation was studied and the H\"older exponents given (for both space and time) are $1/4- \epsilon$. Bertini and Cancrini \cite{BertiniCancrini94Intermittence} stated in their paper that the random field solution for the parabolic Anderson model with initial data satisfying \eqref{E2:BC-InitD} belongs to $C_{\frac{1}{4}-,\frac{1}{2}-}(\mathbb{R}_+^*\times\mathbb{R})$. In \cite{PospisilTribe07Parameter,Shiga94Two}, the authors showed that if the initial data is a continuous function with certain exponentially growing tails, then \begin{align}\label{E2:Shiga} u\in C_{\frac{1}{4}-,\frac{1}{2}-}(\mathbb{R}_+\times\mathbb{R}),\quad\text{a.s.} \end{align} Sanz-Sol\'e and Sarr\`a \cite{SanSoleSarra99Holder} considered the stochastic heat equation over $\mathbb{R}^d$ with spatially homogeneous colored noise which is white in time. Let $\tilde{\mu}$ be the spectral measure satisfying \[ \int_{\mathbb{R}^d}\frac{\tilde{\mu}(\ensuremath{\mathrm{d}} \xi)}{\left(1+|\xi|^2\right)^\eta}<+\infty, \quad\text{for some $\eta\in \;]0,1[$.} \] They proved that if the initial data is a bounded $\rho$-H\"older continuous function for some $\rho\in \;]0,1[$, then the solution is in \[ u(t,x)\in C_{\frac{1}{2}(\rho\wedge (1-\eta))-, \rho\wedge (1-\eta)-} \left(\mathbb{R}_+^*\times\mathbb{R}\right)\;. \] For the case of space-time white noise, the spectral measure $\tilde{\mu}$ is Lebesgue measure and hence $\eta$ can be $1/2-\epsilon$ for any $\epsilon>0$. Their result (\cite[Theorem 4.3]{SanSoleSarra99Path}) reduces to \[ u(t,x)\in C_{\left(\frac{1}{4}\wedge\frac{\rho}{2}\right)-, \left(\frac{1}{2}\wedge\rho\right)-}\left(\mathbb{R}_+^*\times\mathbb{R}\right)\;. \] More recently, Conus {\it et al} proved in their paper \cite[Lemma 9.3]{ConusEct12Initial} that the random field solution is H\"older continuous in $x$ with exponent $1/2-\epsilon$ (for initial data that is a finite measure). They did not give the regularity estimate over the time variable. In their papers \cite{DalangKhNualart07HittingAdditive, DalangKhNualart09HittingMult}, Dalang, Khoshnevisan and Nualart considered a system of heat equations with vanishing initial conditions subject to space-time white noise, and proved that the solution is jointly H\"older continuous with exponents $1/4-$ in time and $1/2-$ in space. We extend the $C_{\frac{1}{4}-,\frac{1}{ 2}-}\left(\mathbb{R}_+^*\times\mathbb{R}\right)$-H\"older continuity result to measure-valued initial data satisfying \eqref{E2:J0finite}. We show that the result in \eqref{E2:Shiga} should exclude the time line $t=0$ unless the initial data $\mu$ is $1/2$-H\"older continuous. The difficulties for the proof of the H\"older continuity of the random field solution lie in the fact that for the initial data satisfying \eqref{E2:J0finite}, the $p$-th moment $\mathbb{E}\left[|u(t,x)|^p\right]$ is neither bounded for $x\in\mathbb{R}$, nor for $t\in [0,T]$. Standard techniques, which isolate the effects of initial data by the $L^p(\Omega)$-boundedness of the solution, fail in our case. Instead, the initial data play an active role in our proof. Note that Fourier transforms are not applicable here because $\mu$ need not be a tempered measure. \section{Main Results}\label{S2:MainRes} Denote the solution to the homogeneous equation \begin{align}\label{E2:Heat-home} \begin{cases} \displaystyle \left(\frac{\partial }{\partial t} - \frac{\nu}{2} \frac{\partial^2 }{\partial x^2}\right) u(t,x) = 0,& x\in \mathbb{R},\; t \in\mathbb{R}_+^*, \\ \displaystyle \quad u(0,\cdot) = \mu(\cdot)\;, \end{cases} \end{align} by \[ J_0(t,x) := \left(\mu* G_{\nu}(t,\cdot)\right)(x) = \int_\mathbb{R} G_\nu(t,x-y)\mu(\ensuremath{\mathrm{d}} y)\:,\quad (t,x)\in\mathbb{R}_+^*\times\mathbb{R}\:. \] \nomenclature{$J_0(t,x)$}{solutions to the homogeneous equation \nomrefpage}\noindent Note that $J_0(t,x)$ is well-defined by the hypothesis \eqref{E2:J0finite}. We formally rewrite the stochastic partial differential equation \eqref{E2:Heat} in the integral form (mild form): \begin{align}\label{E2:WalshSI} u(t,x) &= J_0(t,x) + I(t,x) \end{align} where \begin{align}\label{E2:WalshSI-I} I(t,x):=\iint_{[0,t]\times\mathbb{R}} G_\nu(t-s,x-y) \rho\left( u(s,y) \right) \dot{W}(\ensuremath{\mathrm{d}} s,\ensuremath{\mathrm{d}} y)\:. \end{align} By convention, $I(0,x)=0$. The above stochastic integral is defined in the sense of Walsh \cite{Walsh86,DalangEtc08Minicourse}. \subsection{Notations and Conventions} Assume that the function $\rho:\mathbb{R}\mapsto \mathbb{R}$ is globally Lipschitz continuous with Lipschitz constant $\LIP_\rho>0$. \nomenclature{$\LIP_\rho$}{Lipschitz constant \nomrefpage} We need some growth conditions on $\rho$: Assume that \begin{align}\label{E2:LinGrow} |\rho(x)|^2 \le \Lip_\rho^2 \left(\Vip^2 +|x|^2\right),\qquad \text{for all $x\in\mathbb{R}$}\;, \end{align} for some constants $\Lip_\rho>0$ and $\Vip \ge 0$. \nomenclature{$\Lip_\rho$}{linear growth constant (upper bound) \nomrefeqpage} \nomenclature{$\Vip$}{constant related to the linear growth (upper bound) \nomrefeqpage} Note that $\sqrt{2}\LIP_\rho\le \Lip_\rho$, and the inequality may be strict. In order to bound the second moment from below, we will sometimes assume that for some constants $\lip_\rho>0$ and $\vip \ge 0$, \begin{align}\label{E2:lingrow} |\rho(x)|^2\ge \lip_{\rho}^2\left(\vip^2+|x|^2\right),\qquad \text{for all $x\in\mathbb{R}$}\;. \end{align} \nomenclature{$\lip_\rho$}{linear growth constant (lower bound) \nomrefeqpage}\noindent \nomenclature{$\vip$}{constant related to the linear growth (lower bound) \nomrefeqpage}\noindent We shall also give special attention to the linear case (the parabolic Anderson model) $\rho(u)=\lambda u$ with $\lambda\ne 0$, which is a special case of the following quasi-linear growth condition: for some constant $\vv\ge 0$, \begin{align}\label{E2:qlinear} |\rho(x)|^2= \lambda^2\left(\vv^2+|x|^2\right),\qquad \text{for all $x\in\mathbb{R}$}\;. \end{align} \nomenclature{$\vv$}{constant related to the parabolic Anderson model \nomrefeqpage} We use the convention that $G_\nu(t,\cdot)\equiv 0$ if $t\le 0$. Hence, the integral region in the stochastic integral in \eqref{E2:WalshSI-I} can be written as $\mathbb{R}_+\times\mathbb{R}$. Define a kernel function \begin{align} \label{E2:K} \mathcal{K}\left(t,x;\nu,\lambda\right) := G_{\frac{\nu}{2}}(t,x) \left(\frac{\lambda^2}{\sqrt{4\pi\nu t}}+\frac{\lambda^4}{2\nu} \: e^{\frac{\lambda^4 t}{4\nu}}\Phi\left(\lambda^2 \sqrt{\frac{t}{2\nu}}\right)\right) \;, \end{align} \nomenclature{$\mathcal{K}\left(t,x;\nu,\lambda\right)$}{a kernel function \nomrefeqpage}\noindent for all $(t,x)\in\mathbb{R}_+^*\times \mathbb{R}$, where $\Phi(x)$ is the probability distribution function of the standard normal distribution: \[ \Phi(x) := \int_{-\infty}^x \frac{e^{-y^2/2}}{\sqrt{2\pi}} \ensuremath{\mathrm{d}} y\;. \] \nomenclature{$\Phi(x)$}{the probability distribution function of the standard normal distribution \nomrefpage}\noindent We also use the error function $\ensuremath{\mathrm{erf}}(x):=\frac{2}{\sqrt{\pi}}\int_0^x e^{-y^2}\ensuremath{\mathrm{d}} y$ and its complement $\ensuremath{\mathrm{erfc}}(x):=1-\ensuremath{\mathrm{erf}}(x)$. Clearly, \begin{align*} \Phi(x) = \frac{1}{2}\left(1+\ensuremath{\mathrm{erf}}\left(x/\sqrt{2}\right)\right)\;, \end{align*} \begin{align*} \ensuremath{\mathrm{erf}}(x) &= 2\Phi\left(\sqrt{2}\:x\right)-1,\quad \ensuremath{\mathrm{erfc}}(x) = 2\left(1-\Phi\left(\sqrt{2}\:x\right)\right)\;. \end{align*} \nomenclature{$\ensuremath{\mathrm{erf}}(x)$}{ the error function $\ensuremath{\mathrm{erf}}(x):=\frac{2}{\sqrt{\pi}}\int_0^x e^{-x^2}\ensuremath{\mathrm{d}} x$.\nomrefpage} \nomenclature{$\ensuremath{\mathrm{erfc}}(x)$}{ the complementary error function $\ensuremath{\mathrm{erfc}}(x) = 1-\ensuremath{\mathrm{erf}}(x)$ \nomrefpage} We use $\star$ to denote the simultaneous convolution in both space and time variables. \nomenclature{$\star$}{space-time convolution \nomrefpage} Define another function \begin{align}\label{E2:H} \mathcal{H}(t;\nu,\lambda):=\left(1\star \mathcal{K} \right)(t,x) = 2 e^{\frac{\lambda^4\: t}{4\nu}}\Phi\left(\lambda^2\sqrt{\frac{t}{2\nu}}\right)-1\;. \end{align} Clearly, $\mathcal{K}\left(t,x;\nu,\lambda\right)$ can be written as \[ \mathcal{K}\left(t,x;\nu,\lambda\right)=G_{\nu/2}(t,x) \left( \frac{\lambda^2}{\sqrt{4\pi\nu t}}+ \frac{\lambda^4}{4\nu} \left[\mathcal{H}(t;\nu,\lambda)+1\right] \right) \;. \] We use the following conventions: \begin{align}\label{E2:K-cm} \mathcal{K}(t,x) &:= \mathcal{K}\left(t,x \: ;\: \nu,\lambda\right)\;, \\ \label{E2:upperK} \overline{\mathcal{K}}(t,x) &:= \mathcal{K}\left(t,x \: ;\: \nu,\Lip_\rho\right)\;,\\ \label{E2:lowerK} \underline{\mathcal{K}}(t,x) &:=\mathcal{K}\left(t,x \:;\: \nu,\lip_\rho \right)\;,\\ \label{E2:hatK} \widehat{\mathcal{K}}_p(t,x) &:=\mathcal{K}\left(t,x \: ;\: \nu,a_{p,\Vip}\:z_p\: \Lip_\rho \right)\;,\quad\text{for all $p>2$}\;, \end{align} where $z_p$ is the universal constant in the Burkholder-Davis-Gundy inequality (in particular, $z_2=1$) and $a_{p,\Vip}$ is a constant defined as \begin{align}\label{E2:a_pv} a_{p,\Vip} \: :=\: \begin{cases} 2^{(p-1)/p}& \text{if $\Vip\ne 0,\: p>2$,}\cr \sqrt{2} & \text{if $\Vip =0,\: p>2$,}\cr 1 & \text{if $p=2$}. \end{cases} \end{align} We only need to keep in mind that $a_{p,\Vip}\le 2$. Note that the kernel function $\widehat{\mathcal{K}}_p(t,x)$ implicitly depends on $\Vip$ through $a_{p,\Vip}$ which will be clear from the context. If $p=2$, then $\widehat{\mathcal{K}}_2(t,x) = \overline{\mathcal{K}}(t,x)$. Similarly $\overline{\mathcal{H}}(t)$, $\underline{\mathcal{H}}(t)$ and $\widehat{\mathcal{H}}_p(t)$ denote the kernel functions with $\lambda$ in $\mathcal{H}(t)$ replaced by $\Lip_\rho$, $\lip_\rho$ and $a_{p,\Vip} z_p \Lip_\rho$, respectively. Again $\widehat{\mathcal{H}}_p(t)$ depends on $\Vip$ implicitly which will be clear from the context. Let us set up the filtered probability space. Let \[ \Big\{\: W_t(A):\:A\in\mathcal{B}_b\left(\mathbb{R}\right),t\ge 0 \: \Big\} \] be a space-time white noise defined on a probability space $(\Omega,\mathcal{F},P)$, where $\mathcal{B}_b\left(\mathbb{R}\right)$ is the collection of Borel measurable sets with finite Lebesgue measure. Let $(\mathcal{F}_t,t\ge 0)$ be the standard filtration generated by this space-time white noise. More precisely, let \[ \mathcal{F}_t^0 := \sigma\left(W_s(A):\: 0\le s\le t, A\in\mathcal{B}_b\left(\mathbb{R}\right)\right)\vee \mathcal{N},\quad t\ge 0 \] be the natural filtration augmented by all $P$-null sets in $\mathcal{F}$. Define $\mathcal{F}_t := \mathcal{F}_{t+}^0 = \wedge_{s>t}\mathcal{F}_s^0$ for any $t\ge 0$. In the following, we fix this filtered probability space $\left\{\Omega,\mathcal{F},\{\mathcal{F}_t:t\ge0\},P\right\}$. We use $\Norm{\cdot}_p$ to denote the $L^p(\Omega)$-norm. \nomenclature{$\Norm{\cdot}_p$}{ the $L^p(\Omega)$-norm \nomrefpage} Denote $\Ceil{p}_2:=2\Ceil{p/2}$, which is the smallest even integer greater than or equal to $p$. \nomenclature{$\Ceil{p}_2$}{the smallest even integer greater than or equal to $p$ \nomrefpage} Let $\mathcal{M}(\mathbb{R})$ be the set of locally finite Borel measures over $\mathbb{R}$. \nomenclature{$\mathcal{M}(\mathbb{R})$}{the set of locally finite Borel measures on $\mathbb{R}$ \nomrefpage} Define \begin{align}\label{E2:MGa} \mathcal{M}_G^{\sd}(\mathbb{R}):= \left\{\mu\in\mathcal{M}(\mathbb{R}): \: \int_\mathbb{R} e^{\sd|x|}|\mu| (\ensuremath{\mathrm{d}} x)<+\infty\right\}\;, \quad \sd\ge 0, \end{align} \nomenclature{$\mathcal{M}_G^{\sd}(\mathbb{R})$}{the set of Borel measures on $\mathbb{R}$ that have exponential decays at infinite \nomrefeqpage}\noindent where $|\mu|=\mu_++\mu_-$ is the Jordan decomposition of a measure into two non-negative measures. We use subscript ``$+$'' to denote the subset of non-negative measures. For example, $\mathcal{M}_+(\mathbb{R})$ is the set of non-negative Borel measures over $\mathbb{R}$ and $\mathcal{M}_{G,+}^{\sd}(\mathbb{R}) = \mathcal{M}_G^{\sd}(\mathbb{R}) \cap \mathcal{M}_+(\mathbb{R})$. A random field $Y=\left(Y(t,x):\: (t,x)\in \mathbb{R}_+^*\times \mathbb{R}\right)$ is said to be {\it $L^p(\Omega)$-continuous}, $p\ge 2$, if for all $(t,x) \in \mathbb{R}_+^*\times\mathbb{R}$, \[ \lim_{\left(t',x'\right)\rightarrow (t,x)}\Norm{Y(t,x)-Y\left(t',x'\right)}_p =0\;. \] \subsection{Existence, Uniqueness and Moments} We first give the definition of the random field solution as follows: \begin{definition}\label{D2:Solution} A solution $u=\left(u(t,x):(t,x)\in\mathbb{R}_+^*\times\mathbb{R}\right)$ to \eqref{E2:Heat} (or \eqref{E2:WalshSI}) is called a {\it random field solution} if \begin{enumerate}[(1)] \item $u$ is adapted, i.e., for all $(t,x)\in\mathbb{R}_+^*\times\mathbb{R}$, $u(t,x)$ is $\mathcal{F}_t$-measurable; \item $u$ is jointly measurable with respect to $\mathcal{B}\left(\mathbb{R}_+^*\times\mathbb{R}\right)\times\mathcal{F}$; \item $\left(G_\nu^2 \star \Norm{\rho(u)}_2^2\right)(t,x)<+\infty$ for all $(t,x)\in\mathbb{R}_+^*\times\mathbb{R}$, and the function $(t,x)\mapsto I(t,x)$ mapping from $\mathbb{R}_+^*\times\mathbb{R}$ into $L^2(\Omega)$ is continuous; \item $u$ satisfies \eqref{E2:Heat} (or \eqref{E2:WalshSI}) almost surely, for all $(t,x)\in\mathbb{R}_+^*\times\mathbb{R}$. \end{enumerate} \end{definition} The first main result is stated as follows. \begin{theorem}[Existence, uniqueness, moments] \label{T2:ExUni} Suppose that \begin{enumerate}[(i)] \item the initial data $\mu$ is a signed Borel measure such that \eqref{E2:J0finite} holds; \item the function $\rho$ is Lipschitz continuous such that the linear growth condition \eqref{E2:LinGrow} holds. \end{enumerate} Then the stochastic integral equation \eqref{E2:WalshSI} has a random field solution $u=\{u(t,x): t>0,x\in\mathbb{R}\}$ (note that $t>0$) in the sense of Definition \ref{D2:Solution}. This solution has the following properties: \begin{enumerate}[(1)] \item $u$ is unique (in the sense of versions); \item $(t,x)\mapsto u(t,x)$ is $L^p(\Omega)$-continuous for all integers $p\ge 2$; \item For all even integers $p\ge 2$, the $p$-th moment of the solution $u(t,x)$ satisfies the upper bound \begin{align} \label{E2:SecMom-Up} \Norm{u(t,x)}_p^2 \le \begin{cases} J_0^2(t,x) + \left(J_0^2\star \overline{\mathcal{K}} \right) (t,x) + \Vip^2 \: \overline{\mathcal{H}}(t),& \text{if $p=2$,}\cr \cr 2J_0^2(t,x) + \left(2J_0^2\star \widehat{\mathcal{K}}_p \right) (t,x) + \Vip^2 \: \widehat{\mathcal{H}}_p(t),& \text{if $p>2$,} \end{cases} \end{align} for all $t>0$, $x\in\mathbb{R}$, and the two-point correlation satisfies the upper bound \begin{multline} \label{E2:TP-Up} \mathbb{E}\left[u(t,x)u\left(t,y\right)\right] \\ \le J_0(t,x)J_0\left(t,y\right)+\Lip_\rho^2\int_0^t\ensuremath{\mathrm{d}} s \int_{\mathbb{R}} \overline{f}(s,z) G_\nu(t-s,x-z) G_\nu(t-s,y-z) \ensuremath{\mathrm{d}} z\\ + \frac{\Lip_\rho^2\Vip^2}{\nu}|x-y| \left(\Phi\left(\frac{|x-y|}{\sqrt{2\nu t}}\right)-1\right) + 2 \Lip_\rho^2 \Vip^2 t\: G_{2\nu}(t,x-y)\;, \end{multline} for all $t>0$, $x,y\in\mathbb{R}$, where $\overline{f}(s,z)$ denotes the right hand side of \eqref{E2:SecMom-Up} for $p=2$; \item If $\rho$ satisfies \eqref{E2:lingrow}, then the second moment satisfies the lower bound \begin{align} \label{E2:SecMom-Lower} \Norm{u(t,x)}_2^2 \ge J_0^2(t,x) + \left(J_0^2 \star \underline{\mathcal{K}} \right) (t,x)+ \vip^2\: \underline{\mathcal{H}}(t) \end{align} for all $t>0$, $x\in\mathbb{R}$, and the two-point correlation satisfies the lower bound \begin{multline} \label{E2:TP-Lower} \mathbb{E}\left[u(t,x)u\left(t,y\right)\right] \\ \ge J_0(t,x)J_0\left(t,y\right) +\lip_\rho^2\int_0^t\ensuremath{\mathrm{d}} s \int_{\mathbb{R}} \underline{f}(s,z) G_\nu(t-s,x-z) G_\nu(t-s,y-z) \ensuremath{\mathrm{d}} z\\ + \frac{\lip_\rho^2 \vip^2}{\nu}|x-y| \left(\Phi\left(\frac{|x-y|}{\sqrt{2\nu t}}\right)-1\right) + 2 \lip_\rho^2 \vip^2 t\: G_{2\nu}(t,x-y)\:, \end{multline} for all $t>0$, $x,y\in\mathbb{R}$, where $\underline{f}(s,z)$ denotes the right hand side of \eqref{E2:SecMom-Lower}; \item In particular, for the quasi-linear case $|\rho(u)|^2=\lambda^2 \left(\vv^2+u^2\right)$, the second moment has the explicit expression \begin{align} \label{E2:SecMom} \Norm{u(t,x)}_2^2 = J_0^2(t,x) + \left(J_0^2 \star \mathcal{K} \right) (t,x)+ \vv^2 \: \mathcal{H}(t)\;, \end{align} for all $t>0$, $x\in\mathbb{R}$, and the two-point correlation is given by \begin{multline} \label{E2:TP} \mathbb{E}\left[u(t,x)u\left(t,y\right)\right] \\ = J_0(t,x)J_0\left(t,y\right) +\lambda^2 \int_0^t\ensuremath{\mathrm{d}} s \int_{\mathbb{R}} f(s,z) G_\nu(t-s,x-z) G_\nu(t-s,y-z) \ensuremath{\mathrm{d}} z\\ + \frac{\lambda^2 \vv^2}{\nu}|x-y| \left(\Phi\left(\frac{|x-y|}{\sqrt{2\nu t}}\right)-1\right) + 2 \lambda^2 \vv^2 t\: G_{2\nu}(t,x-y)\:, \end{multline} for all $t>0$, $x,y\in\mathbb{R}$, where $f(s,z)=\Norm{u(s,z)}_2^2$ is defined in \eqref{E2:SecMom}. \end{enumerate} \end{theorem} \begin{corollary}[Lebesgue initial data] \label{C2:TP-Lebesgue} Suppose that $|\rho(u)|^2=\lambda^2(\vv^2+u^2)$ and $\mu$ is Lebesgue measure. Then for all $t>0$ and $x,y\in\mathbb{R}$, \begin{multline} \label{E2:TP-Lebesgue} \mathbb{E}\left[u(t,x)u\left(t,y\right)\right] = 1+(1+\vv^2) \Bigg(\exp\left(\frac{\lambda ^4 t-2 \lambda ^2 |x-y|}{4 \nu }\right) \\ \times \ensuremath{\mathrm{erfc}}\left(\frac{|x-y|-\lambda ^2 t}{2 \sqrt{\nu t}}\right) - \ensuremath{\mathrm{erfc}}\left(\frac{|x-y|}{2 \sqrt{\nu t}}\right) \Bigg)\;. \end{multline} In particular, when $y=x$, we have \begin{align} \label{E2:SecMom-Lebesgue} \mathbb{E}\left[|u(t,x)|^2\right] = 1+(1+\vv^2)\mathcal{H}(t)\;. \end{align} \end{corollary} \begin{remark}\label{R2:TP-Lebesgue} If $\rho(u)=u$ (i.e., $\lambda=1$ and $\vv=0$), then the second moment formula \eqref{E2:SecMom-Lebesgue} recovers, in the case $n=2$, the moment formulae of Bertini and Cancrini \cite[Theorem 2.6]{BertiniCancrini94Intermittence}: \[ \mathbb{E}\left[|u(t,x)|^n\right] = 2 \exp\left\{\frac{n(n^2-1)}{4!\: \nu} t\right\} \Phi\left(\sqrt{ \frac{n(n^2-1)}{12\nu} t}\right). \] As for the two-point correlation function, Bertini and Cancrini \cite[Corollary 2.4]{BertiniCancrini94Intermittence} gave the following integral form: \begin{align} \label{E2:TP-Lebesgue-BC} \mathbb{E}\left[u(t,x)u\left(t,y\right)\right] = \int_0^t \ensuremath{\mathrm{d}} s \frac{|x-y|}{\sqrt{\pi \nu s^3}} \exp\left\{-\frac{(x-y)^2}{4\nu s}+\frac{t-s}{4\nu}\right\}\Phi\left(\sqrt{\frac{t-s}{2\nu}}\right)\:. \end{align} This integral can be evaluated explicitly and equals \begin{align*} \exp\left(\frac{t-2 |x-y|}{4 \nu }\right) \ensuremath{\mathrm{erfc}}\left(\frac{|x-y|-t}{\sqrt{4\nu t}}\right)\:, \end{align*} so their formula differs from \eqref{E2:TP-Lebesgue}. The difference is a term $\ensuremath{\mathrm{erf}}\left(|x-y|/\sqrt{4\nu t}\right)$. By letting $x=y$ in the two-point correlation function, both results do give the correct second moment (the difference term is zero for $x=y$). However, for $x\ne y$, this is not the case. For instance, as $t$ tends to zero, the correlation function should have a limit equal to one, while \eqref{E2:TP-Lebesgue-BC} has limit zero. The argument in \cite{BertiniCancrini94Intermittence} should be modified as follows (we use the notations in their paper): (4.6) on p. 1398 should be \[ \mathbb{E}_{0}^{\beta,1}\left[ \exp\left(\frac{L_t^{\xi}(\beta)}{\sqrt{2\nu}}\right) \right] = \int_0^t P_\xi(\ensuremath{\mathrm{d}} s) \mathbb{E}_0^\beta\left[\exp\left(\frac{L_{t-s}(\beta)}{\sqrt{2\nu}}\right)\right] + P(T_\xi \ge t)\:. \] The extra term is the last term, which is \[ P(T_\xi \ge t) = \int_t^\infty \frac{|\xi|}{\sqrt{2\pi s^3}} \exp\left(-\frac{\xi^2}{2s}\right) \ensuremath{\mathrm{d}} s = \ensuremath{\mathrm{erf}}\left(\frac{|\xi|}{\sqrt{2t}}\right) = \ensuremath{\mathrm{erf}}\left(\frac{\left|x-x'\right|}{\sqrt{4\nu t}}\right)\;. \] With this term, \eqref{E2:TP-Lebesgue} is recovered. \end{remark} \begin{example}[Higher moments for Lebesgue initial data] \label{Ex2:MomLeb} Suppose that $\mu(\ensuremath{\mathrm{d}} x) = \ensuremath{\mathrm{d}} x$. Clearly, $J_0(t,x)\equiv 1$. By the above bound \eqref{E2:SecMom-Up}, we have \begin{align*} \mathbb{E}[|u(t,x)|^p] \le 2^{p-1} + 2^{p/2-1}\left(2+\Vip^2\right)^{p/2} \exp\left\{\frac{a_{p,\Vip}^4 \: z_p^4 \: p \: \Lip_\rho^4\: t}{8\nu}\right\} \left|\Phi\left(a_{p,\Vip}^2 \:\Lip_\rho^2 z_p^2\sqrt{\frac{t}{2\nu}}\right)\right|^{p/2}\:. \end{align*} We can replace $z_p$ by $2\sqrt{p}$, and $a_{p,\Vip}$ by $2$. Then the upper Lyapunov exponent of order $p$ defined in \eqref{E2:Lyapunov} is bounded by \[ \overline{\lambda}_p \le \frac{2^5\: p^3 \Lip_\rho^4}{\nu}\:. \] If $\Vip=0$, we can replace $a_{p,\Vip}$ by $\sqrt{2}$ instead of $2$, which gives a slightly better bound $\overline{\lambda}_p\le 2^3 p^3 \Lip_\rho^4/\nu$. In particular, for the parabolic Anderson model $\rho(u) = \lambda u$, we have \[ \overline{\lambda}_p\le 2^3 p^3 \lambda^4/\nu\;, \] which is consistent with Bertini and Cancrini's formulae $\lambda_p= \frac{\lambda^4}{4! \nu} p(p^2-1)$ (see \cite[(2.40)]{BertiniCancrini94Intermittence}). \end{example} \begin{corollary}[Dirac delta initial data]\label{C2:TP-Delta} Suppose that $|\rho(u)|^2=\lambda^2(\vv^2+u^2)$ and $\mu$ is the Dirac delta measure with a unit mass at zero. Then for all $t>0$ and $x,y\in\mathbb{R}$, \begin{multline} \label{E2:TP-Delta} \mathbb{E}\left[u(t,x)u\left(t,y\right)\right] = G_\nu(t,x)G_\nu\left(t,y\right) -\vv^2 \ensuremath{\mathrm{erfc}}\left(\frac{|x-y|}{2 \sqrt{\nu t}}\right) +\left( \frac{\lambda ^2}{4\nu} G_{\nu/2}\left(t,\frac{x+y}{2}\right)+\vv^2\right)\\ \times \exp\left(\frac{\lambda ^4 t-2 \lambda ^2 |x-y|}{4 \nu }\right) \ensuremath{\mathrm{erfc}}\left(\frac{|x-y|-\lambda ^2 t}{2 \sqrt{\nu t}}\right)\:. \end{multline} In addition, when $y=x$, we have \begin{align} \label{E2:SecMom-Delta} \mathbb{E}\left[|u(t,x)|^2\right] = \frac{1}{\lambda^2}\mathcal{K}(t,x)+\vv^2\mathcal{H}(t)\;. \end{align} \end{corollary} \begin{remark}\label{R2:TP-Delta} If $\rho(u)=u$ (i.e., $\lambda=1$ and $\vv=0$), then the second moment formula \eqref{E2:SecMom-Delta} recovers the result by Bertini and Cancrini \cite[(2.27)]{BertiniCancrini94Intermittence}: \[ \mathbb{E}\left[|u(t,x)|^2\right] = \frac{1}{2\pi\nu t} e^{- \frac{x^2}{\nu t}} \left[ 1+\sqrt{\frac{\pi t}{\nu}} e^{\frac{t}{4\nu}} \Phi\left(\sqrt{\frac{t}{2\nu}}\right) \right], \] which equals $\mathcal{K}\left(t,x;\nu/2,1/\sqrt{4\pi \nu}\right)$. As for the two-point correlation function, Bertini and Cancrini \cite[Corollary 2.5]{BertiniCancrini94Intermittence} gave the following integral form: \begin{multline} \label{E2:TP-Delta-BC} \mathbb{E}\left[u(t,x)u\left(t,y\right)\right] \\ = \frac{1}{2\pi\nu t}\exp\left\{-\frac{x^2+y^2}{2\nu t}\right\} \int_0^1 \frac{|x-y|}{\sqrt{4\pi\nu t}} \frac{1}{\sqrt{s^3(1-s)}} \exp\left\{-\frac{(x-y)^2}{4\nu t}\frac{1-s}{s}\right\} \\ \times\left(1+\sqrt{\frac{\pi t (1-s)}{\nu}} \exp\left\{\frac{t}{2\nu} \frac{1-s}{2}\right\} \Phi\left(\sqrt{\frac{t(1-s)}{2\nu}}\right) \right) \ensuremath{\mathrm{d}} s\;. \end{multline} This integral can be evaluated explicitly, and is equal to \[ = G_\nu(t,x) G_\nu\left(t,y\right) + \frac{1}{4\nu} G_{\frac{\nu}{2}}\left(t,\frac{x+y}{2}\right) \exp\left(\frac{t-2|x-y|}{4\nu}\right) \ensuremath{\mathrm{erfc}}\left(\frac{|x-y|- t }{\sqrt{4\nu t }}\right)\:. \] This coincides with our result \eqref{E2:TP-Delta} for $\vv=0$ and $\lambda=1$. \end{remark} \begin{example}[Higher moments for delta initial data] \label{Ex:GrowthDeta1} Suppose that $\mu = \delta_0$ and $\Vip=0$. Let $p\ge 2$ be an even integer. Clearly, $J_0(t,x)\equiv G_\nu(t,x)$. Then, by \eqref{E2:SecMom-Up}, we have that \begin{align*} \mathbb{E}\left[ |u(t,x)|^p \right] &\le 2^{p-1} G_\nu^{p}(t,x)+2^{(p-2)/2}\left|\left(2 G_\nu^2\star \widehat{\mathcal{K}}_p\right)(t,x)\right|^{p/2}\\ &\le 2^{p-1} G_\nu^{p}(t,x)+2^{(p-2)/2}\Lip_\rho^{-p} z_p^{-p} \left|\widehat{\mathcal{K}}_p(t,x)\right|^{p/2}\\ &= 2^{p-1} G_\nu^{p}(t,x)+ 2^{p-1}G_{\nu/2}^{p/2}(t,x)\left(\frac{1}{\sqrt{4\pi\nu t}}+ \frac{z_p^2\Lip_\rho^2}{\nu}\: e^{\frac{z_p^4\Lip_\rho^4 t}{\nu}}\Phi\left(z_p^2\Lip_\rho^2 \sqrt{\frac{2t}{\nu}}\right)\right)^{p/2}, \end{align*} (the second inequality requires a proof). Hence, for all $x\in\mathbb{R}$, the upper Lyapunov exponent \eqref{E2:Lyapunov} of order $p$ is bounded by \[ \overline{\lambda}_p \le \frac{\Lip_\rho^4 \: z_p^4\: p}{2\nu} \le \frac{2^3\: p^3 \: \Lip_\rho^4 }{\nu}\;, \] where the last inequality is due to the fact that $z_p\le 2\sqrt{p}$ for all $p\ge 2$. Note that this upper bound is identical to the case of Lebesgue initial data. We can also calculate the exponential growth indices explicitly in this case: \[ \lim_{t\rightarrow+\infty}\frac{1}{t} \sup_{|x|> \alpha t}\log\mathbb{E}\left[|u(t,x)|^p\right] \le -\frac{\alpha^2 p}{2\nu} + \frac{\Lip_\rho^4 \: p \: z_p^4}{2\nu}\;,\quad \text{for all $\alpha \ge 0$}\;. \] Hence, the upper growth indices of order $p$ is bounded by $\overline{\lambda}(p) \le z_p^2 \Lip_\rho^2$. Similarly, one can derive that $\underline{\lambda}(2)\ge \lip_\rho^2/2$. Finally, since $\underline{\lambda}(2)\le \underline{\lambda}(p)$ for all $p\ge 2$, we have that, for all even integers $p\ge 2$, \[ \frac{\lip_\rho^2}{2} \le \underline{\lambda}(p) \le \overline{\lambda}(p) \le z_p^2 \Lip_\rho^2 \:. \] Similar bounds are obtained for more general initial data: see Theorem \ref{T2:Growth} below. \end{example} This following proposition shows that initial data cannot be extended beyond measures. \begin{proposition}\label{P2:D-Delta} Suppose that the initial data is $\mu= \delta_0^{'}$, the derivative of the Dirac delta measure at zero. Then the parabolic Anderson model $\rho(u) = \lambda u$ ($\lambda\ne 0$) does not have a random field solution in the sense of Definition \ref{D2:Solution}. \end{proposition} \subsection{Exponential Growth Indices} As an application of the above second moment formula, we partially answer the first open problem proposed by Conus and Khoshnevisan in \cite{ConusKhosh10Farthest}: the limits over $t$ in the definitions of these two indices do exist when $n=2$ and the lower and upper growth indices of order $2$ (see \eqref{E2:GrowInd-0} and \eqref{E2:GrowInd-1}) coincide. Before stating the main result, we first give some explanation concerning the exponential growth indices defined in \eqref{E2:GrowInd-0} and \eqref{E2:GrowInd-1}. When the initial data is localized, for example, when it has compact support, we expect that the position of high peaks of the solution will exhibit a certain wave propagation phenomenon. As shown in Figure \ref{F2:GrowthIndices}, when $\alpha$ is sufficiently large, it is likely that there is no high peaks outside of the space-time cone --- the shaded region. Hence, the limit over $t$ should be negative. The largest $\alpha$ such that this limit remains negative is then defined to be the upper growth index $\overline{\lambda}(p)$. On the other hand, when $\alpha$ is very small, say $\alpha = 0$, then there must be some high peaks in the shaded region so that the limit becomes positive. Hence, the smallest $\alpha$ such that this limit is positive is defined to be the lower growth index $\underline{\lambda}(p)$. \begin{figure}[htbp] \centering \includegraphics[scale=1.2]{GrowthIndices-I} \qquad\qquad \includegraphics[scale=1.2]{GrowthIndices-II} \caption{Illustration of the exponential growth indices. The initial data, depicted by the curve, is localized around the origin. } \label{F2:GrowthIndices} \end{figure} \begin{theorem}[Exponential growth indices] \label{T2:Growth} The following bounds hold: \begin{enumerate}[(1)] \item If $|\rho(u)|^2\le \Lip_\rho^2\left(\Vip^2+u^2\right)$ with $\Vip=0$ (which implies $\vip = \vv = 0$) and the initial data $\mu\in \mathcal{M}_{G}^{\sd}(\mathbb{R})$ for some $\sd > 0$, then for all $p\ge 2$, \[ \bar{\lambda}(p)\le \begin{cases} \displaystyle \frac{\sd \nu}{2} +\frac{z_{\Ceil{p}_2}^4\Lip_\rho^4}{2\nu \sd}\:, & \displaystyle \text{if}\quad 0\le \sd< \frac{z_{\Ceil{p}_2}^2\Lip_\rho^2}{\nu}\:,\cr \displaystyle z_{\Ceil{p}_2}^2\: \Lip_\rho^2\;, & \displaystyle \text{if}\quad \sd \ge \frac{z_{\Ceil{p}_2}^2\Lip_\rho^2}{\nu}\:, \end{cases} \] where $z_m$, $m\in \mathbb{N}$, $m\ge 2$, are the universal constants in the Burkholder-Davis-Gundy inequality. In particular, for $p=2$, \begin{align} \bar{\lambda}(2)\le \begin{cases} \displaystyle \frac{\sd\nu}{2} +\frac{\Lip_\rho^4}{8\nu \sd}\;, & \displaystyle \text{if}\quad 0\le \sd< \frac{\Lip_\rho^2}{2\nu}\;,\cr \displaystyle \frac{1}{2}\Lip_\rho^2\;, & \displaystyle \text{if}\quad \sd \ge \frac{\Lip_\rho^2}{2\nu}\:. \end{cases} \label{E2:Growth-L2} \end{align} \item If $|\rho(u)|^2\ge \lip_\rho^2\left(\vip^2+u^2\right)$ with $\vip=0$, then \[ \underline{\lambda}(p)\ge \frac{\lip_\rho^2}{2},\qquad \text{for all $\mu\in \mathcal{M}_{+}(\mathbb{R})$, $\mu\ne 0$ and all $p\ge 2$}\:; \] otherwise, if $\vip\ne 0$, then \[ \underline{\lambda}(p) = \overline{\lambda}(p)=+\infty,\qquad \text{for all $\mu\in \mathcal{M}_{+}(\mathbb{R})$ and $p\ge 2$}\:; \] \item In particular, for the quasi-linear case $|\rho(u)|^2=\lambda^2\left(\vv^2+u^2\right)$ with $\lambda\ne 0$, if $\vv=0$ and $\sd \ge \frac{\lambda^2}{2\nu}$, then \[ \underline{\lambda}(2)=\bar{\lambda}(2)=\lambda^2/2,\qquad \text{for all $\mu\in \mathcal{M}_{G,+}^{\sd}(\mathbb{R})$, $\mu\ne 0$}\:; \] otherwise, if $\vv\ne 0$, then \[ \underline{\lambda}(p) = \overline{\lambda}(p)=+\infty,\qquad \text{for all $\mu\in \mathcal{M}_{+}(\mathbb{R})$ and $p\ge 2$}\;. \] \end{enumerate} \end{theorem} This theorem generalizes the results by Conus and Khoshnevisan \cite{ConusKhosh10Farthest} in several aspects: (i) more general initial data are allowed; (ii) both non trivial upper bound and lower bounds are given (compare with Theorem 1.1 \cite{ConusKhosh10Farthest}) for the Laplace operator case; (iii) for the parabolic Anderson model, the exact transition is proved (see Theorem 1.3 and the first open problem in \cite{ConusKhosh10Farthest}) for $n=2$ and the Laplace operator case; (iv) our discussions above cover the case $\rho(0)\ne 0$. \begin{example}[Dirac delta initial data] Suppose that $\Vip=\vip=0$. Clearly, $\delta_0 \in \mathcal{M}_{G,+}^{\sd}(\mathbb{R})$ for all $\sd\ge 0$. Hence, the above theorem implies that for all even integers $k\ge 2$, \[ \frac{\lip_\rho^2}{2} \le \underline{\lambda}(k) \le \overline{\lambda}(k) \le z_k^2 \Lip_\rho^2\;. \] This recovers the previous calculation in Example \ref{Ex:GrowthDeta1}. \end{example} \begin{proposition}\label{P2:Example-Exponents} Consider the parabolic Anderson model $\rho(u) =\lambda u$, $\lambda\ne 0$ with the initial data $\mu(\ensuremath{\mathrm{d}} x) = e^{-\sd|x|}\ensuremath{\mathrm{d}} x$ ($\sd>0$). Then we have \[ \underline{\lambda}(2)=\overline{\lambda}(2) = \begin{cases} \displaystyle \frac{\sd\nu}{2} + \frac{\lambda^4}{8 \sd \nu}& \text{if $\;\;\displaystyle 0<\sd\le \frac{\lambda^2}{2\: \nu}$}\;,\cr\cr \displaystyle \frac{\lambda^2}{2} & \text{if $\;\;\displaystyle \sd\ge \frac{\lambda^2}{2\: \nu}$}\;. \end{cases} \] \end{proposition} This proposition shows that for all $\sd\in \left]0,+\infty\right]$, the exact phase transition occurs, and hence our upper bounds \eqref{E2:Growth-L2} in Theorem \ref{T2:Growth} for the upper growth index $\overline{\lambda}(2)$ are sharp. \subsection{Sample Path Regularity} \begin{theorem}\label{T2:Holder} Suppose that $\rho$ is Lipschitz continuous. Then the solution $u(t,x)=J_0(t,x)+I(t,x)$ to \eqref{E2:Heat} has the following sample path regularity: \begin{enumerate}[(1)] \item If the initial data $\mu$ is an $\alpha$-H\"older continuous function ($\alpha\in \;]0,1]$) over $\mathbb{R}$ satisfying \eqref{E2:J0finite}, then \[ J_0\in C_{\frac{1}{2},\alpha} \left(\mathbb{R}_+\times\mathbb{R}\right) \;\cup\; C_{\frac{1}{2},1}\left(\mathbb{R}_+^*\times\mathbb{R}\right)\;, \quad\text{and}\quad I\in C_{\frac{1}{4}-,\frac{1}{2}-}\left(\mathbb{R}_+\times\mathbb{R}\right),\quad \text{a.s.} \] Therefore, \[ u=J_0+I\in C_{\frac{1}{4}-,\left(\frac{1}{2}-\right)\wedge \alpha}\left(\mathbb{R}_+\times\mathbb{R}\right) \;\cup\; C_{\frac{1}{4}-,\frac{1}{2}-}\left(\mathbb{R}_+^*\times\mathbb{R}\right) ,\quad \text{a.s.} \] \item If the initial data $\mu$ is a continuous function satisfying \eqref{E2:J0finite}, then \[ J_0\in C_{\frac{1}{2},1}\left(\mathbb{R}_+^*\times\mathbb{R}\right)\;, \quad\text{and} \quad I \in C_{\frac{1}{4}-,\frac{1}{2}-}\left(\mathbb{R}_+\times\mathbb{R}\right),\quad \text{a.s.} \] Therefore, \[ u=J_0+I\in C_{\frac{1}{4}-,\frac{1}{2}-}\left(\mathbb{R}_+^*\times\mathbb{R}\right) ,\quad \text{a.s.} \] \item If the initial data $\mu$ is a signed Borel measure satisfying \eqref{E2:J0finite}, then \[ J_0\in C_{\frac{1}{2},1}\left(\mathbb{R}_+^*\times\mathbb{R}\right)\;, \quad\text{and}\quad I\in C_{\frac{1}{4}-,\frac{1}{2}-}\left(\mathbb{R}_+^*\times\mathbb{R}\right),\quad \text{a.s.} \] Therefore, \[ u=J_0+I\in C_{\frac{1}{4}-,\frac{1}{2}-}\left(\mathbb{R}_+^*\times\mathbb{R}\right) ,\quad \text{a.s.} \] \end{enumerate} \end{theorem} \begin{example}[Dirac delta initial data]\label{E2:Holder-Delta} Suppose $\rho(u)= \lambda u$ with $\lambda \ne 0$. If $\mu=\delta_0$, then neither $J_0(0,x)$ nor $\lim_{t\rightarrow 0_+} \Norm{I(t,x)}_2$ is continuous in $x$. For $J_0(0,x) = \delta_0(x)$, this is clear. As for $\lim_{t\rightarrow 0_+} \Norm{I(t,x)}_2$, by Corollary \ref{C2:TP-Delta} (with $\vv=0$), we have \[ \Norm{I(t,x)}_2^2= \frac{1}{\lambda^2}\mathcal{K}(t,x) -G_\nu^2(t,x) = \frac{\lambda^2}{2\nu}e^{\frac{\lambda^4 t}{4\nu}}\Phi\left(\lambda^2\sqrt{\frac{t}{2\nu}}\right) G_{\nu/2}(t,x)\;. \] Therefore, \[ \lim_{t\rightarrow 0_+} \Norm{I(t,x)}_2^2 = \begin{cases} 0& \text{if $x\ne 0$}\;,\cr +\infty & \text{if $x=0$}\;. \end{cases} \] \end{example} \bibliographystyle{abbrv}
\section{Introduction}\label{sec1 This article is a continuation of \cite{KOT}. In \cite{KOT}, we have established a Toponogov type triangle comparison theorem (TCT) for a certain class of Finsler manifolds whose radial flag curvatures are bounded below by that of a von Mangoldt surface of revolution (see Theorem~\ref{TCT} for the precise statement). In this article, we prove several applications of our Toponogov theorem on the relationship between the topology and the curvature of a Finsler manifold. We remark that, compared to the Riemannian case, there are only a small number of such kind of results, e.g., Rademacher's quarter pinched sphere theorem (\cite{Ra}), Shen's finiteness theorem under lower Ricci and mean (or ${\mathbf S}$-) curvature bounds (\cite{Shvol}), the second author's generalized splitting theorems under nonnegative weighted Ricci curvature (\cite{Ospl}), and the first author's generalized diameter sphere theorem with radial flag curvature bounded from below by $1$ as an application of TCTs (\cite{K}). \medskip In order to state our results, let us introduce several notions in Finsler geometry as well as the geometry of radial curvature. Let $(M, F, p)$ denote a pair of a forward complete, connected, $n$-dimensional $C^\infty$-Finsler manifold $(M,F)$ with a base point $p \in M$, and $d: M \times M \longrightarrow [0, \infty)$ denote the distance function induced from $F$. We remark that the {\em reversibility} $F(-v)=F(v)$ is not assumed in general, so that $d(x, y) \not= d(y,x)$ is allowed. For a local coordinate $(x^i)^{n}_{i=1}$ of an open subset $\ensuremath{\mathcal{O}} \subset M$, let $(x^i, v^j)_{i,j=1}^{n}$ be the coordinate of the tangent bundle $T\ensuremath{\mathcal{O}}$ over $\ensuremath{\mathcal{O}}$ such that \[ v:= \sum_{j = 1}^{n} v^j \frac{\partial}{\partial x^j}\Big|_{x}, \qquad \ x \in \ensuremath{\mathcal{O}}. \] For each $v \in T_xM \setminus \{0\}$, the positive-definite $n \times n$ matrix \[ \big( g_{ij} (v) \big)_{i,j= 1}^{n}:= \left( \frac{1}{2} \frac{\partial^2 (F^2)}{\partial v^i \partial v^j}(v) \right)_{i, j = 1}^{n} \] provides us the Riemannian structure $g_{v}$ of $T_x M$ by \[ g_{v} \left( \sum_{i = 1}^{n} a^i \frac{\partial}{\partial x^i}\bigg|_{x}, \sum_{j = 1}^{n} b^j \frac{\partial}{\partial x^j}\bigg|_{x} \right) := \sum_{i,j = 1}^{n} g_{ij} (v) a^ib^j. \] This is a Riemannian approximation (up to the second order) of $F$ in the direction $v$. For two linearly independent vectors $v, w \in T_{x} M \setminus \{0\}$, the {\em flag curvature} is defined by \[ K_M (v, w) := \frac{g_{v} (R^{v} (w, v)v, w)}{g_{v} (v, v) g_{v} (w, w) - g_{v} (v, w)^2}, \] where $R^{v}$ denotes the curvature tensor induced from the Chern connection (see \cite[\S 3.9]{BCS} for details). We remark that $K_M (v, w)$ depends not only on the \emph{flag} $\{sv + tw\,|\, s, t \in \ensuremath{\mathbb{R}}\}$, but also on the \emph{flag pole} $\{sv\,|\, s> 0\}$. Given $v,w \in T_xM \setminus \{0\}$, define the \emph{tangent curvature} by \[ \ensuremath{\mathcal{T}}_M(v, w) := g_X\big( D^Y_Y Y(x) - D^X_Y Y(x), X(x) \big), \] where the vector fields $X,Y$ are extensions of $v,w$, and $D_{v}^{w}X(x)$ denotes the covariant derivative of $X$ by $v$ with reference vector $w$. Independence of $\ensuremath{\mathcal{T}}_M(v,w)$ from the choices of $X,Y$ is easily checked. Note that $\ensuremath{\mathcal{T}}_M \equiv 0$ if and only if $M$ is of \emph{Berwald type} (see \cite[Propositions~7.2.2, 10.1.1]{Sh}). In Berwald spaces, for any $x,y \in M$, the tangent spaces $(T_xM, F|_{T_xM})$ and $(T_yM, F|_{T_yM})$ are mutually linearly isometric (cf.~\cite[Chapter~10]{BCS}). In this sense, $\ensuremath{\mathcal{T}}_M$ measures the variety of tangent Minkowski normed spaces. \medskip Let $\widetilde{M}$ be a complete $2$-dimensional Riemannian manifold, which is homeomorphic to $\ensuremath{\mathbb{R}}^{2}$. Fix a base point $\tilde{p} \in \widetilde{M}$. Then, we call the pair $(\widetilde{M}, \tilde{p})$ a {\em model surface of revolution} if its Riemannian metric $d\tilde{s}^2$ is expressed in terms of the geodesic polar coordinate around $\tilde{p}$ as \[ d\tilde{s}^2 = dt^2 + f(t)^2d \theta^2, \qquad (t,\theta) \in (0,\infty) \times \ensuremath{\mathbb{S}}_{\tilde{p}}^1, \] where $f : (0, \infty) \longrightarrow \ensuremath{\mathbb{R}}$ is a positive smooth function which is extensible to a smooth odd function around $0$, and $\ensuremath{\mathbb{S}}^{1}_{\tilde{p}} := \{ v \in T_{\tilde{p}} \widetilde{M} \,|\, \| v \| = 1 \}$. Define the {\em radial curvature function} $G: [0,\infty) \longrightarrow \ensuremath{\mathbb{R}}$ such that $G(t)$ is the Gaussian curvature at $\tilde{\gamma}(t)$, where $\tilde{\gamma}:[0,\infty) \longrightarrow \widetilde{M}$ is any (unit speed) meridian emanating from $\tilde{p}$. Note that $f$ satisfies the differential equation $f''+Gf=0$ with initial conditions $f(0) = 0$ and $f'(0) = 1$. We call $(\widetilde{M}, \tilde{p})$ a {\em von Mangoldt surface} if $G$ is non-increasing on $[0,\infty)$. Paraboloids and $2$-sheeted hyperboloids are typical examples of von Mangoldt surfaces. An atypical example of such a surface is the following. \begin{example}{\rm (\cite[Example 1.2]{KT1})} Set $f (t) := e^{- t^{2}} \tanh t$ on $[0, \infty)$. Then the non-compact surface of revolution $(\widetilde{M}, \tilde{p})$ with $d\tilde{s}^2 = dt^2 + f(t)^2d \theta^2$ is of von Mangoldt type, and $G$ changes the sign. Indeed, $\lim_{t \downarrow 0}G (t) = 8$ and $\lim_{t \to \infty}G (t) = - \infty$. \end{example} We say that a Finsler manifold $(M, F, p)$ has the {\em radial flag curvature bounded below by that of a model surface of revolution $(\widetilde{M}, \tilde{p})$} if, along every unit speed minimal geodesic $\gamma: [0,l) \longrightarrow M$ emanating from $p$, we have \[ K_{M} \big(\dot{\gamma}(t), w \big) \ge G (t) \] for all $t \in [0, l)$ and $w \in T_{\gamma(t)}M$ linearly independent to $\dot{\gamma}(t)$. \bigskip We set \begin{equation}\label{G_p} \mathcal{G}_p (x) := \{ \dot{\gamma}(l) \in T_{x}M \,|\, \text{$\gamma$ is a minimal geodesic segment from $p$ to $x$} \}, \end{equation} where $\gamma:[0,l] \longrightarrow M$ with $l=d(p,x)$, and \[ B_r^+ (p):=\{ x \in M \,|\, d(p,x)<r \}, \qquad \mathop{\mathrm{diam}}\nolimits\left( \partial B_r^+ (p) \right):=\sup_{q_1,\,q_2 \in \partial B_r^+ (p)}d(q_1,q_2). \] Then, our first main result is a finiteness theorem of topological type. \begin{thmA}\label{2012_03_26_main1} Let $(M, F, p)$ be a forward complete, non-compact, connected $C^{\infty}$-Finsler manifold whose radial flag curvature is bounded below by that of a von Mangoldt surface $(\widetilde{M}, \tilde{p})$ satisfying $f'(\rho) =0$ and $G(\rho) \ne0$ for a unique $\rho \in (0, \infty)$. Assume that, for some $t_0 > \rho$, \begin{enumerate}[{\rm (1)}] \item $\mathop{\mathrm{diam}}\nolimits(\partial B_t^+ (p)) = O (t^\alpha)$ for some $\alpha \in (0, 1)$ as $t \to \infty$, \item $g_v(w, w) \ge F(w)^2$ for all $x \in M \setminus \overline{B_{t_0}^+ (p)}$, $v \in \ensuremath{\mathcal{G}}_p (x)$ and $w \in T_xM$, \item $\ensuremath{\mathcal{T}}_{M} (v, w) = 0$ for all $x \in M \setminus \overline{B_{t_0}^+ (p)}$, $v \in \ensuremath{\mathcal{G}}_p (x)$ and $w \in T_xM$, \item the reverse curve $\bar{c}(s):=c(l-s)$ of any minimal geodesic segment $c:[0,l] \longrightarrow M \setminus \overline{B_{t_0}^+ (p)}$ is geodesic. \end{enumerate} Then $M$ has finite topological type, i.e., $M$ is homeomorphic to the interior of a compact manifold with boundary. \end{thmA} \begin{remark} All conditions in Theorem A are sufficient ones that make our TCT hold (see Theorem~\ref{TCT}): The condition (1) guarantees the condition (1) in Theorem~\ref{TCT}. The biggest obstruction when we establish a TCT in Finsler geometry is the covariant derivative even though $F$ is reversible. By the condition (2) and $f' <0$ on $(\rho, \infty)$ (because of $f'(\rho) =0$ and $G(\rho) \ne0$), we can overcome the obstruction, i.e., thanks to the (2), we can transplant the strictly concaveness of $\widetilde{M} \setminus \overline{B_{t_0} (\tilde{p})}$ to $M \setminus \overline{B_{t_0}^+ (p)}$ (see \cite[Section 3]{KOT} for more details), where the convexity on $\widetilde{M} \setminus \overline{B_{t_0} (\tilde{p})}$ arises from the negative second fundamental form for $f' <0$ on $(\rho, \infty)$. Note that the (2) is the \emph{$2$-uniform convexity} with the sharp constant (see \cite{Ouni}), but, in our situation, {\bf only for special} points and directions. This means that the convexity holds only along all minimal geodesic segments emanating from $p$ in our theorem. It is very natural thing to assume that the condition (3), if we employ a {\bf Riemannian} model surface of revolution $\widetilde{M}$ as a reference surface. Here note that $\ensuremath{\mathcal{T}}_{\widetilde{M}} \equiv 0$. It is {\bf not difficult} to construct non-Riemannian spaces satisfying (2) and (3) (see Example \ref{exa2013_09_03}). \end{remark} \begin{example}(\cite{KOT})\label{exa2013_09_03}\\[1mm] $\bullet$ Let $(M,g, p)$ be a complete non-compact Riemannian manifold whose radial (sectional) curvature is bounded below by that of a von Mangoldt surface $(\widetilde{M}, \tilde{p})$ satisfying $f'(\rho) =0$ and $G(\rho) \ne0$ for unique $\rho \in (0, \infty)$. Modify (the unit spheres of) $g$ on $M \setminus B^+_{\rho}(p)$, outside a neighborhood of $\bigcup_{z \in M \setminus B^+_{\rho}(p)} \ensuremath{\mathcal{G}}_p(z)$, in such a way that the (2) holds. Note that the resulting non-Riemannian metric still satisfies the (3), because this modification does not affect $g_v$ for $v \in \bigcup_{z \in M \setminus B^+_{\rho}(p)} \ensuremath{\mathcal{G}}_p(z)$.\\[2mm] $\bullet$ Let $(M, F, p)$ be the Finsler manifold satisfying the radial flag curvature conditions on Theorem A. If $F$ is Riemann on $M \setminus \overline{B_\rho^+(p)}$, then $(M, F, p)$ satisfies all conditions in Theorem A except for the (1). E.g., \[ F(v) = \begin{cases} \ \sqrt{g(v, v)} + \beta (v) \ &\text{on} \ B_\rho^+(p)\\[4mm] \ \sqrt{g(v, v)} \ &\text{on} \ M \setminus \overline{B_\rho^+(p)} \end{cases} \] etc. \end{example} By changing the structure of $F$, we can reduce a few conditions in Theorem A: \begin{corollary} Let $(M, F, p)$ be a forward complete, non-compact, connected $C^{\infty}$-Finsler manifold whose radial flag curvature is bounded below by that of a von Mangoldt surface $(\widetilde{M}, \tilde{p})$ satisfying $f'(\rho) =0$ and $G(\rho) \ne0$ for unique $\rho \in (0, \infty)$. Assume that, for some $t_0 > \rho$, \begin{enumerate}[{\rm (1)}] \item $\mathop{\mathrm{diam}}\nolimits(\partial B_t^+ (p)) = O (t^\alpha)$ for some $\alpha \in (0, 1)$ as $t \to \infty$, \item $g_v(w, w) \ge F(w)^2$ for all $x \in M \setminus \overline{B_{t_0}^+ (p)}$, $v \in \ensuremath{\mathcal{G}}_p (x)$ and $w \in T_xM$, \end{enumerate} If $F$ is of {\bf Berwald} type on $M \setminus \overline{B_{t_0}^+ (p)}$, then $M$ has finite topological type. \end{corollary} \begin{remark}The condition (4) in Theorem A always holds, if $F$ is reversible on $M \setminus \overline{B_{t_0}^+ (p)}$. In the case where $F$ is Riemannian, the diameter growth bound (1) seems to be very restrictive. Indeed, if we employ a non-negatively curved non-compact model surface of revolution $(\widetilde{M}, \tilde{p})$ having the diameter growth $o(t^{1/ 2})$, then $M$ is isometric to the $n$-dimensional model space $\widetilde{M}^n$ (see \cite[Theorem~1.2]{ST}, \cite[Example~1.1]{KT2}). Hence, if $F$ is Riemannian, then we can prove, without the growth condition, the finiteness of topological type of a complete non-compact Riemannian manifold with radial curvature bounded below by that of an {\bf arbitrary} non-compact model surface of revolution admitting a finite total curvature(see \cite[Theorem~2.2]{KT2} and \cite[Theorem~1.3]{TK}). \end{remark} By an entirely different technique, if $F$ is reversible, then we can improve Theorem~A as follows: \begin{thmB} Let $(M, F, p)$ be a forward complete, non-compact, connected $C^{\infty}$-Finsler manifold whose radial flag curvature is bounded below by that of a von Mangoldt surface $(\widetilde{M}, \tilde{p})$ satisfying $f'(\rho) =0$ and $G(\rho) \ne0$ for unique $\rho \in (0, \infty)$. Assume that, for some $t_0 > \rho$, \begin{enumerate}[{\rm (1)}] \item $\mathop{\mathrm{diam}}\nolimits(\partial B_t^+ (p)) = O (t^\alpha)$ for some $\alpha \in (0, 1)$ as $t \to \infty$, \item $g_v(w, w) \ge F(w)^2$ for all $x \in M \setminus \overline{B_{t_0}^+ (p)}$, $v \in \ensuremath{\mathcal{G}}_p (x)$ and $w \in T_xM$, \item $\ensuremath{\mathcal{T}}_{M} (v, w) = 0$ for all $x \in M \setminus \overline{B_{t_0}^+ (p)}$, $v \in \ensuremath{\mathcal{G}}_p (x)$ and $w \in T_xM$. \end{enumerate} If $F$ is reversible, then $M$ is diffeomorphic to $\ensuremath{\mathbb{R}}^n$ and, for every unit speed minimal geodesic $\gamma: [0,\infty) \longrightarrow M$ emanating from $p$, we have $K_{M}(\dot{\gamma}(t), w) = G (t)$ for all $t>0$. \end{thmB} \begin{remark} In Theorem B, we can remove the condition (3), if we additionally assume that $M$ is of Berwald type. The result related to Theorem B is Shiohama and the third author's \cite[Theorem~1.2]{ST}, where they proved that a complete non-compact Riemannian manifold is isometric to the $n$-dimensional model space $\widetilde{M}^n$ if its radial curvature is bounded below by that of a non-compact model surface of revolution $\widetilde{M}$ satisfying $\int_1^{\infty} f(t)^{-2} \,dt = \infty$. Observe that our von Mangoldt surface always satisfies this integration assumption. However, in our Finsler situation, it is difficult (and in fact impossible in many cases) to obtain isometry to a model space. That is, spaces of constant flag curvatures are not unique. E.g., all Minkowski normed spaces have the flat flag curvature and all Hilbert geometries satisfy $K_M \equiv -1$ (cf.\ \cite{Shspr}). Other result related to Theorem~B is the first and the third authors' \cite[Theorem~1.1]{KT3} on a complete non-compact connected Riemannian manifold with smooth convex boundary. \end{remark} \section{A Toponogov type triangle comparison theorem}\label{sec2 We first recall the Toponogov type triangle comparison theorem established in \cite[Theorem~1.2]{KOT}. We refer to \cite{BCS} and \cite{Sh} for the basics of Finsler geometry. Let $(M,F,p)$ be a forward complete, connected $C^\infty$-Finsler manifold with a base point $p \in M$, and denote by $d$ its distance function. The forward completeness guarantees that any two points in $M$ can be joined by a minimal geodesic segment (by the Hopf-Rinow theorem, \cite[Theorem~6.6.1]{BCS}). Since $d(x,y) \neq d(y,x)$ in general, we also introduce \[ d_{\rm{m}} (x, y) := \max\{d(x, y), d(y, x)\}. \] It is clear that $d_{\rm{m}}$ is a distance function of $M$. We can define the `angles' with respect to $d_{\rm{m}}$ as follows. \begin{definition}{\bf (Angles)}\label{2012_03_24_def2.3} Let $c :[0,a] \longrightarrow M$ be a unit speed minimal geodesic segment (i.e., $F(\dot{c}) \equiv 1$) with $p \not\in c([0,a])$. The \emph{forward} and the \emph{backward angles} $\overrightarrow{\angle}(pc(s)c(a))$, $\overleftarrow{\angle}(pc(s)c(0)) \in [0,\pi]$ at $c(s)$ are defined via \begin{align*} \cos \overrightarrow{\angle}\big( pc(s)c(a) \big) &:= -\lim_{h \downarrow 0} \frac{d(p,c(s+h)) -d(p, c(s))}{d_{\rm{m}} (c(s), c(s + h))} \quad \text{for $s \in [0,a)$}, \\ \cos \overleftarrow{\angle}\big( pc(s)c(0) \big) &:= \lim_{h \downarrow 0} \frac{d(p, c(s)) - d (p, c(s-h))}{d_{\rm{m}} (c(s -h), c(s))} \quad \text{for $s \in (0, a]$}. \end{align*} (These limits indeed exist in $[-1,1]$ thanks to the definition of $d_{\rm m}$, see \cite[Lemma~2.2]{KOT}). \end{definition} \begin{definition}{\bf (Forward triangles)}\label{def2.2_ft} For three distinct points $p, x, y \in M$, \[ \triangle (\overrightarrow{px}, \overrightarrow{py}) := (p, x, y; \gamma, \sigma, c) \] will denote the \emph{forward triangle} consisting of unit speed minimal geodesic segments $\gamma$ emanating from $p$ to $x$, $\sigma$ from $p$ to $y$, and $c$ from $x$ to $y$. Then the corresponding \emph{interior angles} $\overrightarrow{\angle}x, \overleftarrow{\angle}y$ at the vertices $x$, $y$ are defined by \[ \overrightarrow{\angle}x := \overrightarrow{\angle}\big( p c(0) c(d(x,y)) \big), \qquad \overleftarrow{\angle}y := \overleftarrow{\angle}\big( p c(d(x,y)) c(0) \big). \] \end{definition} \begin{definition}{\bf (Comparison triangles)} Fix a model surface of revolution $(\widetilde{M}, \tilde{p})$. Given a forward triangle $\triangle (\overrightarrow{px}, \overrightarrow{py})= (p, x, y; \gamma, \sigma, c)\subset M$, a geodesic triangle $\triangle (\tilde{p}\tilde{x} \tilde{y}) \subset \widetilde{M}$ is called its \emph{comparison triangle} if \[ \tilde{d}(\tilde{p}, \tilde{x}) = d(p, x), \qquad \tilde{d}(\tilde{p},\tilde{y}) = d(p, y), \qquad \tilde{d}(\tilde{x},\tilde{y}) = L_{\rm{m}}(c) \] hold, where we set \[ L_{\rm{m}}(c):= \int^{d(x,\,y)}_0 \max\{ F(\dot{c}),F(-\dot{c}) \} \,ds. \] \end{definition} Now, the main result of \cite{KOT} asserts the following. \begin{theorem}[TCT, \cite{KOT}]\label{TCT} Assume that $(M, F, p)$ is a forward complete, connected $C^{\infty}$-Finsler manifold whose radial flag curvature is bounded below by that of a von Mangoldt surface $(\widetilde{M}, \tilde{p})$ satisfying $f'(\rho) =0$ and $G(\rho) \ne0$ for a unique $\rho \in (0, \infty)$. Let $\triangle (\overrightarrow{px}, \overrightarrow{py}) = (p, x, y; \gamma, \sigma, c) \subset M$ be a forward triangle satisfying that, for some open neighborhood $\ensuremath{\mathcal{N}}(c)$ of $c$, \begin{enumerate}[$(1)$] \item $c([0,d(x,y)]) \subset M \setminus \overline{B^+_{\rho} (p)}$, \item $g_v(w,w) \ge F(w)^2$ for all $z \in \ensuremath{\mathcal{N}}(c)$, $v \in \ensuremath{\mathcal{G}}_p(z)$ and $w \in T_zM$, \item $\ensuremath{\mathcal{T}}_{M}(v,w) = 0$ for all $z \in \ensuremath{\mathcal{N}}(c)$, $v \in \ensuremath{\mathcal{G}}_p(z)$ and $w \in T_zM$, and the reverse curve $\bar{c}(s):=c(d(x,y)-s)$ of $c$ is also geodesic. \end{enumerate} If such $\triangle (\overrightarrow{px}, \overrightarrow{py})$ admits a comparison triangle $\triangle (\tilde{p}\tilde{x} \tilde{y})$ in $\widetilde{M}$, then we have $\overrightarrow{\angle} x \ge \angle \tilde{x}$ and $\overleftarrow{\angle} y \ge \angle \tilde{y}$. \end{theorem} \begin{remark} If a von Mangoldt surface $(\widetilde{M}, \tilde{p})$ satisfies $G(\rho) = 0$ for a unique $\rho \in (0, \infty)$, then $f'(\rho) =0$ and $f'(t) >0$ on $(\rho, \infty)$. In this case, Theorem \ref{TCT} holds, if $F(w)^2 \ge g_v(w,w)$ for all $z \in \ensuremath{\mathcal{N}}(c)$, $v \in \ensuremath{\mathcal{G}}_p(z)$ and $w \in T_zM$ as in (2). For this, see \cite[Remark 2.10]{K} \end{remark} \section{Fundamental tools on model surfaces}\label{sec2.5 We next introduce some fundamental tools in the geometry of model surfaces of revolution. We refer to \cite[Chapter 7]{SST} for more details. Let $(\widetilde{M}, \tilde{p})$ be a non-compact model surface of revolution with its metric $d\tilde{s}^2 = dt^2 + f(t)^2d \theta^2$ on $(0,a) \times \ensuremath{\mathbb{S}}_{\tilde{p}}^1$. Given a unit speed geodesic $\tilde{c} : [0, a) \longrightarrow \widetilde{M}$ $(0 < a \le \infty)$ expressed as $\tilde{c}(s) = (t(s), \theta(s))$, there exists a non-negative constant $\nu$ such that \begin{equation}\label{2012_03_28_clairaut} \nu = f\big( t(s) \big)^2 |\theta' (s)| = f\big( t(s) \big) \sin \angle\big( \dot{\tilde{c}}(s), (\partial/\partial t)|_{\tilde{c}(s)} \big) \end{equation} for all $s \in [0,a)$. The equation (\ref{2012_03_28_clairaut}) is called the {\em Clairaut relation}, and $\nu$ is called the {\em Clairaut constant} of $\tilde{c}$. Note that $\nu=0$ if and only if $\tilde{c}$ is (a part of) a meridian. Since $\tilde{c}$ has unit speed, we deduce from $|t'|^2+|f(t)\theta'|^2=1$ that \[ |t'(s)| = \frac{\sqrt{f(t(s))^{2} - \nu^2}}{f(t(s))}. \] Thus we observe that $t'(s) = 0$ if and only if $f(t(s)) = \nu$. Moreover, if $a<\infty$, then the length $L(\tilde{c})$ of $\tilde{c}$ is not less than \begin{equation}\label{length-ineq_c} t(a) - t(0) + \frac{\nu^{2}}{2} \int_{t(0)}^{t(a)} \frac{1}{f(t)\sqrt{f(t)^{2} -\nu^{2}}} \,dt. \end{equation} The proof of \eqref{length-ineq_c} can be found in (the proof of) \cite[Lemma~2.1]{ST}. \section{Proof of Theorem~A}\label{sec3 Let $(M,F,p)$, $f$ and $\rho$ be as in Theorem~A. The following fact on the cut loci of a von Mangoldt surface is important. \begin{remark}\label{2012_09_19_rem4.1} The cut locus $\mathop{\mathrm{Cut}}\nolimits (\tilde{x})$ of $\tilde{x} \not= \tilde{p}$ is either an empty set, or a ray properly contained in the meridian $\theta^{-1} (\theta (\tilde{x}) + \pi)$ opposite to $\tilde{x}$. Moreover, the endpoint of $\mathop{\mathrm{Cut}}\nolimits(\tilde{x})$ is the first conjugate point to $\tilde{x}$ along the minimal geodesic from $\tilde{x}$ passing through $\tilde{p}$ (\cite[Main Theorem]{T}). \end{remark} We first show an auxiliary lemma on the model surface. \begin{lemma}\label{2012_03_26_lem3.1} If two distinct points $\tilde{x},\tilde{y} \in \widetilde{M} \setminus \overline{B_{\rho}(\tilde{p})}$ satisfy $\tilde{d}(\tilde{p}, \tilde{x}) \le \tilde{d}(\tilde{p}, \tilde{y})$, then \[ \angle \big( \dot{\tilde{c}} (0), (\partial / \partial t)|_{\tilde{x}} \big) <\pi /2 \] holds for any unit speed minimal geodesic segment $\tilde{c}$ emanating from $\tilde{x}$ to $\tilde{y}$. In particular, we have $\tilde{c}([0, d(\tilde{x}, \tilde{y})]) \subset \widetilde{M} \setminus \overline{B_\rho (\tilde{p})}$. \end{lemma} \begin{proof} Let us write $\tilde{c}(s) = (t(s), \theta (s))$. Suppose that $\angle (\dot{\tilde{c}} (0), (\partial / \partial t)|_{\tilde{x}}) \ge \pi/2$ which is equivalent to $t'(0) \le 0$. Since $f' < 0$ on $(\rho, \infty)$ because of $f'(\rho) =0$ and $G(\rho) \ne0$ for a unique $\rho \in (0, \infty)$, it follows from \cite[(7.1.15)]{SST} that \[ t''(0) = f\big( t(0) \big) f'\big( t(0) \big) \theta'\big( t(0) \big)^2 <0. \] Hence $t(s)$ is decreasing on $[0, \delta]$ for some small $\delta > 0$. Since $t(d(\tilde{x}, \tilde{y})) = \tilde{d}(\tilde{p}, \tilde{y}) \ge \tilde{d}(\tilde{p}, \tilde{x}) = t(0)$, there exists $s_0 \in (0, \tilde{d}(\tilde{p}, \tilde{y}))$ such that $t'(s_0) = 0$ and $t(s_0)<t(0)$. By the Clairaut relation \eqref{2012_03_28_clairaut}, for any $s \in [0, \tilde{d}(\tilde{p}, \tilde{y})]$, we observe \[ f \big( t(s_0) \big) = f\big( t(s) \big) \sin \angle \big( \dot{\tilde{c}} (s), (\partial / \partial t)|_{\tilde{c}(s)} \big) \le f\big( t(s) \big). \] Since $f'<0$ on $(\rho,\infty)$ and $t(s_0)<t(0)$, this shows $t(s_0)<\rho$. Thus $\tilde{c}$ intersects the parallel $t= \rho$ twice in $\theta^{-1} ((\theta (\tilde{x}), \theta (\tilde{x}) + \pi))$, where we assume that $\theta (\tilde{x}) \le \theta (\tilde{y})$. However, since $f'(\rho) =0$, the parallel $t = \rho$ is geodesic. Therefore (by rotation) $\tilde{x}$ has a cut point in $\theta^{-1}((\theta (\tilde{x}), \theta (\tilde{x}) + \pi))$. This contradicts the structure of $\mathop{\mathrm{Cut}}\nolimits (\tilde{x})$ (see Remark \ref{2012_09_19_rem4.1}). $\hfill \Box$ \end{proof} \begin{lemma}\label{2012_03_29_lem3.2} If two points $x, y \in M \setminus \overline{B_{\rho}^+ (p)}$ satisfy $d(p, y) > d(p, x) \gg t_0$, then \[ c \big( [0,d(x, y)] \big) \cap \partial B_{t_0}^+ (p) = \emptyset \] holds for any minimal geodesic segment $c$ emanating from $x$ to $y$, where $t_0 > \rho$ is as in the assumption of Theorem~{\rm A}. \end{lemma} \begin{proof} By the assumption (1) of Theorem~A, there is a constant $C > 0$ such that \begin{equation}\label{2012_03_29_lem3.2_1} \frac{\mathop{\mathrm{diam}}\nolimits(\partial B_t^+ (p))}{t^\alpha} < C \end{equation} for all $t \gg t_0$. Suppose that $c ([0, d(x, y)]) \cap \partial B_{t_0}^+ (p) \not= \emptyset$ for some minimal geodesic segment $c$ emanating from $x$ to $y$. Let $S$ be the set of all $s \in (0, d(x, y))$ such that $c(s) \in \partial B_{t_0}^+ (p)$, and set $s_0:= \sup S$. Since $d(p, y) > d(p, x)$, there exists $s_1 \in (s_0, d(x, y))$ such that $c(s_1) \in \partial B_{t_1}^+(p)$, where $t_1:= d(p, x)$. Observe from the triangle inequality that \[ s_1 -s_0 =d\big( c(s_0), c(s_1) \big) \ge d\big( p, c(s_1) \big) -d\big( p, c(s_0) \big) =t_1 -t_0. \] Since $\mathop{\mathrm{diam}}\nolimits(\partial B_{t_1}^+ (p)) \ge s_1 > s_1 -s_0 \ge t_1 -t_0$, we obtain \[ \frac{\mathop{\mathrm{diam}}\nolimits(\partial B_{t_1}^+ (p))}{t_1^\alpha} > t_1^{1-\alpha} - \frac{t_0}{t_1^\alpha}. \] This contradicts (\ref{2012_03_29_lem3.2_1}), because $t_1 \gg t_0$ and $\alpha<1$. $\hfill \Box$ \end{proof} Analogously to \cite{GS}, we define critical points of the distance function $d_p:=d(p,\cdot)$ as follows. Recall \eqref{G_p} for the definition of $\ensuremath{\mathcal{G}}_p (x)$. \begin{definition}\label{2012_03_31_def3.3} We say that a point $x \in M$ is a {\em forward critical point} for $p \in M$ if, for every $w \in T_{x} M \setminus \{0\}$, there exists $v \in \ensuremath{\mathcal{G}}_p (x)$ such that $g_v (v, w) \le 0$. \end{definition} An important consequence of the criticality is that, for any $y \in M$ and any forward triangle $\triangle (\overrightarrow{px}, \overrightarrow{py})$, we have $\overrightarrow{\angle}x \le \pi/2$. We can prove Gromov's isotopy lemma \cite{G} by a similar arguments to the Riemannian case. \begin{lemma}\label{2012_03_31_lem3.4} Given $0 < r_1 < r_2 \le \infty$, if $\overline{B_{r_2}^+(p)} \setminus B_{r_1}^+(p)$ has no critical point for $p \in M$, then $\overline{B_{r_2}^+(p)} \setminus B_{r_1}^+(p)$ is homeomorphic to $\partial B_{r_1}^+(p) \times [r_1, r_2 ]$. \end{lemma} Now we are ready to prove Theorem~A. \medskip \noindent {\it Proof of Theorem~{\rm A}.} By virtue of Lemma~\ref{2012_03_31_lem3.4}, it is sufficient to prove that the set of forward critical points for $p$ is bounded. Suppose that there is a divergent sequence $\{ q_{i} \}_{i \in \ensuremath{\mathbb{N}}}$ of forward critical points for $p$. Then there exist $i_1, i_2 \in \ensuremath{\mathbb{N}}$ such that \[ d(p, q_{i_2}) > d(p, q_{i_1}) \gg t_0 > \rho. \] Let $c:[0, a] \longrightarrow M$ be a minimal geodesic segment emanating from $q_{i_1}$ to $q_{i_2}$. Note that $\overrightarrow{\angle} (p c(0) c(a)) \le \pi/2$ by the criticality of $q_{i_1}$, and $c([0, a]) \cap \partial B_{t_0}^+ (p) = \emptyset$ by Lemma~\ref{2012_03_29_lem3.2}. We first consider the case where $d(p,q_{i_1}) =\min_{s \in [0,a]} d(p,c(s))$. For sufficiently small $s_1\in (0,a)$, the forward triangle $\triangle (\overrightarrow{pq_{i_1}},\overrightarrow{pc(s_1)})$ admits a comparison triangle $\triangle (\tilde{p} \widetilde{q_{i_1}} \widetilde{c(s_1)})$ in $\widetilde{M}$. Then, by Theorem~\ref{TCT}, we observe that $\angle\widetilde{q_{i_1}} \le \overrightarrow{\angle} (p c(0) c(a)) \le \pi/2$. Since \[ \tilde{d}(\tilde{p},\widetilde{q_{i_1}}) =d(p, q_{i_1}) \le d\big( p, c(s_1) \big) =\tilde{d}\big( \tilde{p}, \widetilde{c(s_1)} \big), \] this contradicts Lemma~\ref{2012_03_26_lem3.1}. If $\min_{s \in [0,a]} d(p,c(s)) <d(p,q_{i_1})$, then we fix $s_0 \in (0,a)$ such that $d(p, c(s_0)) = \min_{s \in [0,a]} d(p, c(s))$. By construction, it holds $\overrightarrow{\angle}(pc(s_0)c(a)) =\pi/2$ (note that $\overrightarrow{\angle}(pc(s_0)c(a)) >\pi/2$ can not happen by Theorem~\ref{TCT}). Thus we derive a contradiction from the same argument as the first case. $\hfill \Box$ \section{Proof of Theorem B}\label{sec4 Let $(M,F,p)$, $f$, $\rho$ and $t_0$ be as in Theorem~B. Suppose that the cut locus $\mathop{\mathrm{Cut}}\nolimits (p)$ of $p$ is not empty. Then, since $M$ is non-compact, $\mathop{\mathrm{Cut}}\nolimits(p)$ is an unbounded set (consider a sequence in the open set $D_p:=\{ v \in U_pM \,|\, \gamma_v((0,\infty)) \cap \mathop{\mathrm{Cut}}\nolimits(p) \neq \emptyset \}$ whose limit belongs to the complement $D_p^c=\{ v \in U_pM \,|\, \gamma_v\ \text{is a ray}\}$, where $U_pM:=T_pM \cap F^{-1}(1)$ and $\gamma_v(t):=\exp_p(tv)$ for $t \ge 0$). Let $N(p)$ denote the set of all points $x \in M$ admitting at least two minimal geodesic segments emanating from $p$ to $x$. Note that $N(p)$ is dense in $\mathop{\mathrm{Cut}}\nolimits (p)$ (see \cite[Proposition 2.6]{TS}). Take a divergent sequence $\{ x_{i} \}_{i \in \ensuremath{\mathbb{N}}} \subset N(p)$, and fix $i_0 \in \ensuremath{\mathbb{N}}$ such that $d(p, x_{i_0}) > t_0$. Since $M$ is non-compact and complete, there exists a unit speed ray $\sigma:[0,\infty) \longrightarrow M$ emanating from $p$. Take a divergent sequence $\{r_j\}_{j \in \ensuremath{\mathbb{N}}} \subset (d(p,x_{i_0}),\infty)$ and, for each $j$, let $c_j:[0, a_j] \longrightarrow M$ be a unit speed minimal geodesic segment emanating from $x_{i_0}$ to $\sigma (r_j)$. By Lemma~\ref{2012_03_29_lem3.2}, $c_j([0,a_j]) \cap \partial B_{t_0}(p) = \emptyset$ holds for all $j \in \ensuremath{\mathbb{N}}$. Take a subdivision $s_0 := 0 < s_1 < \cdots < s_{k -1} <s_k := a_j$ of $[0, a_j]$ such that $\triangle (\overrightarrow{p c_j (s_{l-1})}, \overrightarrow{p c_j (s_l)})$ admits a comparison triangle $\widetilde{\triangle}^l := \triangle (\tilde{p} \widetilde{c_j (s_{l-1})} \widetilde{c_j (s_l)}) \subset \widetilde{M}$ for each $l = 1,2, \ldots, k$. Note that, by the reversibility of $F$, \begin{equation}\label{2012_04_01_thm4.1_2} \tilde{d}(\widetilde{c_j (s_{l-1})}, \widetilde{c_j (s_l)}) = L_{\rm{m}}(c_j |_{[s_{l-1},\,s_l]}) = s_l - s_{l-1}. \end{equation} It follows from Theorem~\ref{TCT} that \begin{equation}\label{2012_04_01_thm4.1_3} \overrightarrow{\angle} c_j (s_{l-1}) \ge \angle\big( \tilde{p} \widetilde{c_j (s_{l-1})} \widetilde{c_j (s_l)} \big), \qquad \overleftarrow{\angle}c_j (s_l) \ge \angle \big( \tilde{p} \widetilde{c_j (s_l)} \widetilde{c_j (s_{l-1})} \big) \end{equation} for each $l = 1,2, \ldots, k$. Starting from $\widetilde{\triangle}^1$, we inductively draw a geodesic triangle $\widetilde{\triangle}^{l + 1} \subset \widetilde{M}$ which is adjacent to $\widetilde{\triangle}^l$ so as to have a common side $\tilde{p} \widetilde{c_j (s_l)}$, where $0 \le \theta (\widetilde{c_j (s_0)}) \le \theta (\widetilde{c_j (s_1)}) \le \cdots \le \theta (\widetilde{c_j (s_k)})$. We observe from the definition of the angles that $\overleftarrow{\angle} c_j (s_l) + \overrightarrow{\angle} c_j (s_l) \le \pi$ for each $l = 1,2, \ldots, k-1$. Together with \eqref{2012_04_01_thm4.1_3}, we obtain \begin{equation}\label{2012_04_01_thm4.1_4} \angle\big( \tilde{p} \widetilde{c_j (s_l)} \widetilde{c_j (s_{l-1})} \big) + \angle\big( \tilde{p} \widetilde{c_j (s_l)} \widetilde{c_j (s_{l+1})} \big) \le \pi. \end{equation} Let $\widehat{\xi}_j:[0, a_j] \longrightarrow \widetilde{M}$ denote the broken geodesic segment consisting of minimal geodesic segments from $\widetilde{c_j (s_{l-1})}$ to $\widetilde{c_j (s_l)}$, $l=1,2,\ldots,k$. We set $\widehat{\xi}_j (s) = (t(\widehat{\xi}_j (s)), \theta (\widehat{\xi}_j (s)))$. Then \eqref{2012_04_01_thm4.1_4} gives us the unit speed (not necessarily minimal) geodesic $\widetilde{\eta}_j :[0, b_j] \longrightarrow \widetilde{M}$ emanating from $\widetilde{c_j (0)}$ to $\widetilde{c_j (a_j)}$ and passing under $\widehat{\xi}_j ([0,a_j])$, i.e., $\theta (\widetilde{\eta}_j) \in [\theta (\widetilde{c_j (0)}), \theta (\widetilde{c_j (a_j)})]$ on $[0, b_j]$ and $t(\widehat{\xi}_j (s)) > t(\widetilde{\eta}_j (b))$ for all $(s,b) \in (0,a_j) \times (0, b_j)$ with $\theta (\widehat{\xi}_j (s)) = \theta (\widetilde{\eta}_j (b))$. On the one hand, by \eqref{2012_04_01_thm4.1_2}, we have \[ L(\widetilde{\eta}_j) \le L(\widehat{\xi}_j) = \sum_{l=1}^k \tilde{d}\big( \widetilde{c_j (s_{l-1})}, \widetilde{c_j (s_l)} \big) = s_k - s_0 = a_j, \] where $L(\widetilde{\eta}_j)$ denotes the length of $\widetilde{\eta}_j$. Moreover, the reversibility of $F$ and the triangle inequality show \begin{equation}\label{2012_04_01_thm4.1_5} L(\widetilde{\eta}_j) \le a_j = d\big( x_{i_0}, \sigma(r_j) \big) \le d(p, x_{i_0}) + r_j . \end{equation} On the other hand, it follows from \eqref{length-ineq_c} that \begin{align*} L(\widetilde{\eta}_j) &\ge r_j - d(p, x_{i_0}) + \frac{\nu_j^{2}}{2} \int_{d(p, \,x_{i_0})}^{r_j} \frac{1}{f(t)\sqrt{f(t)^2 - \nu_j^{2}}} \,dt\\ & \ge r_j - d(p, x_{i_0}) + \frac{\nu_j^{2}}{2} \int_{d(p, \,x_{i_0})}^{r_j} f(t)^{-2} \,dt, \end{align*} where $\nu_j$ denotes the Clairaut constant of $\widetilde{\eta}_j$. Together with \eqref{2012_04_01_thm4.1_5}, we find \[ 4 d(p, x_{i_0}) \ge \nu_j^2 \int_{d(p, \,x_{i_0})}^{r_j} f(t)^{-2} \,dt. \] Since $f$ is decreasing on $(\rho, \infty)$ because of $f'(\rho) =0$ and $G(\rho) =0$ for a unique $\rho \in (0, \infty)$, this implies $\lim_{j \to \infty} \nu_j = 0$. Hence we have \[ \lim_{j \to \infty} \angle\big( \dot{\widetilde{\eta}}_j(0), (\partial / \partial t)|_{\widetilde{\eta}_j(0)} \big) =0. \] Combining this with $\angle(\dot{\widetilde{\eta}}_j(0),(\partial/\partial t)|_{\widetilde{\eta}_j(0)}) = \pi-\angle(\tilde{p} \widetilde{c_j(0)} \widetilde{c_j(s_1)})$ and \eqref{2012_04_01_thm4.1_3}, we obtain $\lim_{j \to \infty} \overrightarrow{\angle} c_j (0)= \pi$. This is a contradiction, since $c_j(0) = x_{i_0}\in N(p)$. Hence $\mathop{\mathrm{Cut}}\nolimits (p)=\emptyset$, so that $M$ is diffeomorphic to $\ensuremath{\mathbb{R}}^n$. The curvature equality follows from the same argument as \cite[Theorem~4.8]{KT3}. $\hfill \Box$
\section{Introduction} The \emph{Kepler} space mission has recently announced the discovery of the hot Jupiter Kepler-17b. It orbits a G2V star with $K_{p}\footnote{\emph{Kepler} magnitude (cf. Sect.~\ref{observations}).}=14.14$ in $P=1.486$~d \citep[][ hereafter D11]{Desertetal11}. Thanks to additional spectra acquired with the SOPHIE spectrograph at the Observatoire de Haute Provence, \citet{Bonomoetal12} (hereafter B12) slightly refined the characterization of the Kepler-17 planetary system. In particular, they found the host star to be slightly hotter and younger than D11, with an effective temperature of $5781 \pm 85$~K and an age $< 1.8$~Gyr. The hosting star Kepler-17 is remarkably active, as shown by the out-of-transit variations in the \emph{Kepler} light curve with a peak-to-peak amplitude of $\approx 4\%$. They are produced by the rotational modulation of active regions, starspots and faculae, on the stellar disc. By means of a periodogram analysis of the light curve, D11 estimated the stellar rotation period to be $P_{\rm rot}=11.89 \pm 0.15$~d. Intriguingly, this value of $P_{\rm rot}$ is an integer multiple of the orbital period, specifically eight times. This could be just a coincidence or the result of a star-planet interaction, as pointed out by D11. In addition, the planet occults starspots during transits, giving rise to the typical bumps observed in the bottom of the transit profiles \citep[e.g., ][]{Silvavalioetal10,SilvavalioLanza11}. The integer ratio between $P_{\rm rot}$ and $P$ allowed D11 to see a ``stroboscopic" effect with the short-cadence \emph{Kepler} data: the spots are ``mapped" by the planet each 45~\ensuremath{^\circ} in longitude. The study of these spot-crossing events reveals that the planet's orbit is prograde and the projected spin-orbit angle is smaller than $10^{\circ}-15^{\circ}$ (see D11 for more details). Kepler-17 is therefore an excellent candidate to study the magnetic activity in a solar-like star younger than the Sun, through the modelling of its out-of-transit flux variations. A similar analysis has been performed for several CoRoT planet-hosting stars. One of them is CoRoT-2, an active G7V star with a rotation period of 4.5~d and an age younger than 0.5~Gyr \citep{Alonsoetal08}. The spot-modelling of its light curve revealed two active longitudes and a cyclic oscillation of the total spotted area with a period of $28.3 \pm 4.3$~d, similarly to the solar Rieger cycles \citep{Lanzaetal09a}. Investigating the magnetic activity of G dwarfs with different ages would in principle allow one to study the Sun in time. Recently, \citet{Frascaetal11} and \citet{Frohlichetal12} have applied a spot-modelling to \emph{Kepler} photometric time series to study three young G- and K-type dwarf stars whose ages range between $\approx 50$ and $\approx 200$~Myr, deriving information on their spot evolution and surface differential rotation. Owing to the faintness and low projected rotation velocity of Kepler-17 ($ v \sin i \approx 5$~km~s$^{-1}$), an investigation of its magnetic activity through Doppler imaging is not feasible. Therefore, modelling the stellar variability and studying the distortions of the transit profile when the planet occults starspots are the only techniques that allow us to derive active longitudes where spots preferentially form, determine the lifetime of active regions and activity complexes, estimate a minimum amplitude for the stellar differential rotation, and possibly discover short-term and/or long-term activity cycles, depending on the length of the time series. \section{Observations} \label{observations} The \emph{Kepler} space telescope has an aperture of 95~cm and is designed to yield nearly continuous, high-precision photometry in the passband $423-897$~nm for $\approx 150, 000$ stars in a fixed field of view to search for planetary transits \citep[see, e.g., ][]{Boruckietal10,Kochetal10}. It orbits the Sun on an Earth-trailing orbit and, to keep the solar arrays pointed towards the Sun, a re-orientation of the spacecraft is required every $\approx 90$ days, a time interval called ``a quarter" in the \emph{Kepler} jargon. The re-orientation of the telescope produces an offset in the photon counting for a given star because its photometric mask is re-defined on a different CCD in the focal plane. The light curve of Kepler-17, publicly available at the MAST archive\footnote{http://archive.stsci.edu/kepler/data\_search/search.php}, covers more than fifteen months of photometric measurements, from 2009 May 13 to 2010 August 23. Observations are distributed along the six quarters Q1-Q6. The raw data with the long-cadence temporal sampling, i.e. one point each 29.42~min, was used for our work. Short-cadence data (one point per minute) available for the last three quarters (Q4-Q6) are not particularly useful for our purpose because we are interested in modelling the out-of-transit variations on the timescale of stellar rotation or longer. Because of the flux offsets between adjacent quarters and the long-term instrumental trends within each quarter (see \citealt{Jenkinsetal10a} and Fig.~1 in D11 for our specific case), the light curves corresponding to different quarters were separately treated as follows. First, planetary transits were removed from each quarter. The flux contamination due to starfield crowding, as estimated by the \emph{Kepler} team, was subtracted from the median value of the flux. Steep variations after the safe modes \citep{Jenkinsetal10b} were removed and long-term trends of clear instrumental origin were corrected by fitting a parabola. Lastly, the flux of each quarter was normalized to its median value then nearly matched the endpoints of adjacent quarters. The final light curve obtained by combining the six quarters contains 20\,924 data points and is shown in Sect.~\ref{light_curve_model} (see Fig.~\ref{lc_bestfit}). The median of the errors of the single photometric measurements is $2.18 \times 10^{-4}$ in relative flux units. \section{Spot modelling of wide-band light curves} \label{spotmodel} Reconstructing the surface brightness distribution from the rotational modulation of the stellar flux is an ill-posed problem, because the variation of the flux vs. rotational phase contains information only on the distribution of the brightness inhomogeneities vs. longitude. The integration over the stellar disc effectively cancels any latitudinal information, particularly when the inclination of the rotation axis along the line of sight is close to $90^{\circ}$, as in the present case \citep[see Sect.~\ref{model_param} and ][]{Lanzaetal09a}. Therefore, we need to include a priori information in the light curve inversion process to obtain a unique and stable map. This is done by computing a maximum entropy (hereinafter ME) map, which has been proven to best reproduce active region distribution and area variations for the Sun \citep[cf. ][]{Lanzaetal07}. For a different modelling approach, based on discrete starspots, see, e.g., \citet{Mosseretal09}. A comparison with other modelling approaches is given in, e.g., \citet{Frohlichetal09,Huberetal10,SilvavalioLanza11}. In the present model, the stellar surface is subdivided into elements, i.e., into 200 squares of side $18^{\circ}$, with each element containing unperturbed photosphere, dark spots, and facular areas. The fraction of the $k$-th element covered by dark spots is indicated by its filling factor $f_{k}$, the fractional area of the faculae is $Qf_{k}$, and the fractional area of the unperturbed photosphere is $1-(Q+1)f_{k}$. The contribution to the stellar flux coming from the $k$-th surface element at time $t_{j}$, where $j=1,..., N$ is an index numbering the $N$ points along the light curve, is given by \begin{eqnarray} \Delta F_{kj} & = & I_{0}(\mu_{kj}) \left\{ 1-(Q+1)f_{k} + c_{\rm s} f_{k} + \right. \nonumber \\ & & \left. Q f_{k} [1+c_{\rm f} (1 -\mu_{kj})] \right\} A_{k} \mu_{kj} {w}(\mu_{kj}), \label{delta_flux} \end{eqnarray} where $I_{0}$ is the specific intensity in the continuum of the unperturbed photosphere at the isophotal wavelength of the observations, $c_{\rm s}$ and $c_{\rm f}$ are the spot and facular contrasts, respectively \citep[cf. ][]{Lanzaetal04}, $A_{k}$ is the area of the $k$-th surface element, \begin{equation} {w} (\mu_{kj}) = \left\{ \begin{array}{ll} 1 & \mbox{if $\mu_{kj} \geq 0$} \\ 0 & \mbox{if $\mu_{kj} < 0$ } \end{array} \right. \end{equation} is its visibility, and \begin{equation} \mu_{kj} \equiv \cos \psi_{kj} = \sin i \sin \theta_{k} \cos [\ell_{k} + \Omega (t_{j}-t_{0})] + \cos i \cos \theta_{k}, \label{mu} \end{equation} is the cosine of the angle $\psi_{kj}$ between the normal to the surface element and the direction of the observer, with $i$ being the inclination of the stellar rotation axis to the line of sight, $\theta_{k}$ the colatitude and $\ell_{k}$ the longitude of the $k$-th surface element, $\Omega$ the angular velocity of rotation of the star ($\Omega \equiv 2 \pi / P_{\rm rot}$), and $t_{0}$ the initial time. The specific intensity in the continuum varies according to a quadratic limb-darkening law, as adopted by \citet{Lanzaetal03} for the Sun, viz. $I_{0} \propto a_{\rm p} + b_{\rm p} \mu + c_{\rm p} \mu^{2}$. The stellar flux computed at time $t_{j}$ is then: $F(t_{j}) = \sum_{k} \Delta F_{kj}$. To warrant a relative precision of about $10^{-5}$ in the computation of the flux $F$, each surface element is further subdivided into elements of $1^{\circ} \times 1^{\circ}$ and their contributions, calculated according to Eq.~(\ref{delta_flux}), are summed up at each given time to compute the contribution of the $18^{\circ} \times 18^{\circ}$ surface element to which they belong. We fitted the light curve by changing the value of the spot-filling factor $f$ over the surface of the star while $Q$ was held constant. Even fixing the rotation period, the inclination, and the spot and facular contrasts \citep[see ][ for details]{Lanzaetal07}, the model has 200 free parameters and suffers from non-uniqueness and instability. To find a unique and stable spot map, we applied ME regularization, as described in \citet{Lanzaetal07}, by minimizing a functional $Z$, which is a linear combination of the $\chi^{2}$ and the entropy functional $S$; i.e., \begin{equation} Z = \chi^{2} ({\vec f}) - \lambda S ({\vec f}), \end{equation} where ${\vec f}$ is the vector of the filling factors of the surface elements, $\lambda > 0$ a Lagrangian multiplier determining the trade-off between light curve fitting and regularization, and { the expression of $S$ is \begin{equation} S = -\sum_{k} w_{k} \left[ f_{k} \log \frac{f_{k}}{m} + (1-f_{k}) \log \frac{1-f_{\rm k}}{1-m} \right], \end{equation} where $w_{k}$ is the relative area of the $k$-th surface element (total surface area of the star $=1$) and $m$ the default spot-covering factor that fixes the limiting values for $f_{k}$ as: $m < f_{k} < (1-m)$. In our modelling we adopted $m=10^{-6}$ \citep[cf. ][]{Lanzaetal98}. } The entropy functional $S$ is constructed in such a way that it attains its maximum value when the star is virtually immaculate ($f_{k} = m$ for every $k$). Therefore, by increasing the Lagrangian multiplier $\lambda$, the weight of $S$ in the model is increased and the area of the spots is progressively reduced. This gives rise to systematically negative residuals between the observations and the best-fit model when $\lambda > 0$. The optimal value of $\lambda$ depends on the information content of the light curve, which in turn depends on the ratio of the amplitude of its rotational modulation to the average standard deviation of its points. In the case of Kepler-17, the amplitude of the rotational modulation is $\approx 0.044$, while the nominal standard deviation of the points is $\approx 2.18 \times 10^{-4}$ in relative flux units (see Sect.~\ref{observations}), giving a signal-to-noise ratio of $\approx 200$. To fix the optimal value of the Lagrangian multiplier $\lambda$, we compared the modulus of the mean of the residuals of the regularized best fit $|\mu_{\rm reg}|$ with the standard error of the residuals themselves, i.e., $\epsilon_{0} \equiv \sigma_{0}/\sqrt{N}$, where $\sigma_{0}$ is the standard deviation of the residuals of the unregularized best fit and $N$ the number of points in each fitted subset of the light curve having a duration $\Delta t_{\rm f}$ (see below). We iterated until $|\mu_{\rm reg}| \simeq \epsilon_{0}$ \citep[cf. the case of \object{CoRoT-2} in ][]{Lanzaetal09a}. For the Sun, by assuming a fixed distribution of the filling factor, it is possible to obtain a good fit of the irradiance changes only for a limited time interval $\Delta t_{\rm f}$, not exceeding 14 days, which is the lifetime of the largest sunspot groups dominating the irradiance variation \citep{Lanzaetal03}. For other active stars, the value of $\Delta t_{\rm f}$ must be determined from the observations themselves, looking for the maximum data extension that allows us a good fit with the applied model (see Sect.~\ref{model_param}). The optimal values of the spot and facular contrasts and of the facular-to-spotted area ratio $Q$ in stellar active regions are unknown a priori. In our model the facular contrast $c_{\rm f}$ and the parameter $Q$ enter as the product $c_{\rm f} Q$, so we can fix $c_{\rm f}$ and vary $Q$, estimating its best value by minimizing the $\chi^{2}$ of the model, as shown in Sect.~\ref{model_param}. Since the number of free parameters of the ME model is large, we used the model of \citet{Lanzaetal03} to fix the value of $Q$. It fits the light curve by assuming only three active regions to model the rotational modulation of the flux plus a uniformly distributed background to account for the variations of the mean light level. This procedure is the same as adopted for, e.g., \object{CoRoT-2} and \object{CoRoT-4} to fix the value of $Q$ \citep[cf. ][]{Lanzaetal09a,Lanzaetal09b}. We assumed an inclination of the rotation axis of Kepler-17 of $ i = 87.2^{\circ}$ (see Sect.~\ref{model_param}). Since the information on spot latitudes that can be extracted from the rotational modulation of the flux for such a high inclination is negligible, the ME regularization virtually puts all spots at the sub-observer latitude (i.e., $90^{\circ} -i \approx 3^{\circ}$) to minimize their area and maximize the entropy. Therefore, we are limited to mapping only the distribution of the active regions vs. longitude, which can be achieved with a resolution of at least $\approx 40^{\circ}-50^{\circ}$ \citep[cf. ][]{Lanzaetal07,Lanzaetal09a}. Our ignorance of the true facular contribution to the light modulation may lead to systematic errors in the active region longitudes derived by our model, as discussed by \citet{Lanzaetal07} for the Sun. \section{Model parameters} \label{model_param} The basic stellar parameters are taken from D11 and B12 and are listed in Table~\ref{model_param_table}. The limb-darkening parameters in the \emph{Kepler} bandpass were derived from the model of the transit as discussed in D11. The rotation period for our spot modelling was initially fixed at exactly eight orbital periods of the planet, following D11, who found that there were spots occulted by the planet that reappeared at the same phase after that time interval, i.e., $11.89$~d. However, we found that the migration rates of the other spots revealed by the modelling of the out-of-transit photometry was minimized by assuming a slightly different rotation period, i.e., $12.01$~d, which we adopted for our modelling. \begin{table} \noindent \caption{Parameters adopted for the modelling of the light curve of Kepler-17.} \centering \begin{tabular}{lrr} \hline & & \\ Parameter & Value & Ref.$^{a}$\\ & & \\ \hline & & \\ Star mass ($M_{\odot}$) & 1.16 & B12 \\ Star radius ($R_{\odot}$) & 1.05 & B12 \\ $T_{\rm eff}$ (K) & 5780 & B12 \\ $\log g$ (cm s$^{-2}$) & 4.53 & B12\\ $a_{\rm p}$ & 0.333 & BL12 \\ $b_{\rm p}$ & 0.929 & BL12 \\ $c_{\rm p}$ & -0.262 & BL12 \\ $P_{\rm rot}$ (days) & 12.01 & BL12 \\ $\epsilon_{\rm rot}$ & $4.66 \times 10^{-5}$ & BL12 \\ Inclination (deg) & 87.22 & D11 \\ $c_{\rm s}$ & 0.677 & L04 \\ $c_{\rm f}$ & 0.115 & L04 \\ $Q$ & 1.6 & BL12 \\ $\Delta t_{\rm f}$ (days) & 8.733 & BL12 \\ & & \\ \hline \label{model_param_table} \end{tabular} ~\\ $^{a}$ References: B12: \citet{Bonomoetal12}; D11: \citet{Desertetal11}; L04: \citet{Lanzaetal04}; BL12: present study. \end{table} The polar flattening of the star owing to the centrifugal potential was computed in the Roche approximation with a rotation period of 12.01~d. The relative difference between the equatorial and the polar radii is $\epsilon_{\rm rot} = 4.66 \times 10^{-5}$, which induces a completely negligible relative flux variation of $\approx 2 \times 10^{-6}$ for a spot coverage of $\approx 4$ percent, as a consequence of the gravity darkening of the equatorial regions of the star. The inclination of the stellar rotation axis is constrained by the observation of D11 that the light anomalies detected in successive transits are compatible with the same spots being repeatedly occulted as they move over the stellar disc owing to stellar rotation. This is possible only if the disc-projected trajectory of the spots remains inside the belt occulted by the planet. In other words, this is an indication that the sky-projected misalignment between the stellar spin and the orbital angular momentum is smaller than $\pm (10^{\circ}-15^{\circ})$ (cf. D11 for details). Although the possibility that the stellar spin is not aligned with the orbital angular momentum cannot be excluded with certainty because only the sky-projected angle between the two vectors has been measured, we regard this occurrence as highly improbable and assume that the rotation axis is inclined with respect to the line of sight of the same angle as the orbit normal. Moreover, the stellar $v \sin i = 4.7 \pm 1.0$~km~s$^{-1}$ and the estimated stellar radius provide an inclination close to $90^{\circ}$ for the stellar rotation period (cf. D11). The maximum time interval $\Delta t_{\rm f}$ that our model can accurately fit with a fixed distribution of active regions was determined by dividing the total interval, $T= 497.785$~d, into $N_{\rm f}$ equal segments, i.e., $\Delta t_{\rm f} = T/N_{\rm f}$, and looking for the minimum value of $N_{\rm f}$ that allows us a good fit of the light curve, as measured by the $\chi^{2}$ statistics. We found that for $N_{\rm f} < 57 $ the quality of the best fit degrades appreciably with respect to higher values, owing to a substantial evolution of the pattern of surface brightness inhomogeneities. Therefore, we adopted $N_{\rm f} = 57$, so that $\Delta t_{\rm f} = 8.733$~d is the maximum time interval to be fitted with a fixed distribution of surface active regions to estimate the best value of the parameter $Q$ (see below). To evaluate the spot contrast, we adopted the same mean temperature difference as derived for sunspot groups from their bolometric contrast, i.e. 560~K \citep{Chapmanetal94}. The effective temperature of the unspotted photosphere is $5780 \pm 85$~K, i.e., very similar to that of the Sun (cf. B12). In other words, we assumed a spot effective temperature of $ 5220$~K, yielding a contrast $c_{\rm s} = 0.677$ in the bolometric passband \citep[cf. ][]{Lanzaetal07}. A different spot contrast changes the absolute spot coverages, but neither significantly affects their longitudes or their time evolution, as discussed in detail by \citet{Lanzaetal09a}. Therefore, since the spot temperature is only estimated, we neglected the difference in the contrast $c_{\rm s}$ between the \emph{Kepler} bandpass and the bolometric passband in our analysis. In our model, the facular contrast is assumed to be solar-like with $c_{\rm f} = 0.115$ \citep{Lanzaetal04}. The best value of the area ratio $Q$ between the faculae and the spots in the active regions was estimated by means of the three-spot model by \citet[][ cf. Sect.~\ref{spotmodel}]{Lanzaetal03}. In Fig.~\ref{qratio}, we plot the ratio $\chi^{2}/ \chi^{2}_{\rm min}$ of the total $\chi^{2}$ of the composite best fit of the entire time series to its minimum value $\chi^{2}_{\rm min}$, versus $Q$, and indicate the 95 percent confidence level as derived from the F-statistics \citep[e.g., ][]{Lamptonetal76}. The choice of $\Delta t_{\rm f} = 8.733$~d allows us to fit the rotational modulation of the active regions for the longest time interval during which they remain stable, modelling both the flux increase due to the facular component when an active region is close to the limb and the flux decrease due to the dark spots when the same region transits across the central meridian of the disc. In this way, a measure of the relative facular and spot contributions can be obtained, leading to an estimate of $Q$. Fig.~\ref{qratio} shows that the best value is $Q=1.6$, with an acceptable range extending from~$\approx 0.6$~to~$\approx 2.6$. Therefore, we adopted $Q=1.6$ for our modelling in Sect.~\ref{results}. We comment on the value of $Q$ in more detail in Sect.~\ref{conclusions}. \begin{figure}[b] \centerline{ \includegraphics[width=8cm,height=6cm,angle=0]{qfac.ps}} \caption{ Ratio of the $\chi^{2}$ of the composite best fit of the entire time series of Kepler-17 to its minimum value vs. the parameter $Q$, i.e., the ratio of the area of the faculae to that of the cool spots in active regions. The horizontal dashed line indicates the 95 percent confidence level for $\chi^{2}/\chi_{\rm min}^{2}$, determining the interval of acceptable $Q$ values. } \label{qratio} \end{figure} \begin{figure*}[!t] \centerline{ \includegraphics[width=12cm,height=16cm,angle=90]{gen_fit_final.ps}} \caption{{\it Upper panel:} The ME-regularized composite best fit to the out-of-transit light curve of Kepler-17 obtained for $Q=1.6$. The flux is in relative units, i.e., normalized to the maximum observed flux along the light curve. {\it Lower panel:} The residuals from the composite best fit versus the time. } \label{lc_bestfit} \end{figure*} \begin{figure}[!b] \centerline{ \includegraphics[width=6cm,height=8cm,angle=90]{hist_res.ps}} \caption{Distribution of the residuals to the composite ME-regularized best fit shown in Fig.~\ref{lc_bestfit}. The dashed line is a Gaussian fit to the residual distribution, while the vertical dotted line marks the zero value of the residuals. } \label{hist_residuals} \end{figure} \section{Results} \label{results} \subsection{Light curve models} \label{light_curve_model} We applied the model of Sect.~\ref{spotmodel} to the out-of-transit light curve of Kepler-17, considering time intervals $\Delta t_{\rm f} = 8.733$~d. The best fit without regularization ($\lambda = 0$) has a mean $\mu_{\rm res} = 3.845 \times 10^{-6}$ and a standard deviation of the residuals $\sigma_{0} = 3.032 \times 10^{-4}$ in relative flux units. The Lagrangian multiplier $\lambda$ is iteratively adjusted until the mean of the residuals $\mu_{\rm res} = -1.583 \times 10^{-5} \simeq \sigma_{0}/\sqrt{N}$, where $N = 367$ is the mean number of points in each fitted light curve interval $\Delta t_{\rm f}$; the standard deviation of the residuals of the regularized best fit is $\sigma = 3.311 \times 10^{-4}$. The composite best fit to the entire light curve is shown in the upper panel of Fig.~\ref{lc_bestfit}, while the residuals are plotted in the lower panel. The best fit is always very good, with a standard deviation of the residuals $\approx 1.52$ times the median of the errors of the photometric measurements as given by the \emph{Kepler} pipeline. The distribution of the residuals is plotted in Fig.~\ref{hist_residuals} and is well fitted by a Gaussian with a standard deviation of $3.056 \times 10^{-4}$ for absolute values of the residuals lower than $ \approx 6 \times 10^{-4}$ in relative flux units. For residuals greater than $\approx 6 \times 10^{-4}$ there is a remarkable asymmetry in the distribution with an excess of positive residuals. A periodogram of the residuals shows a highly significant peak at a period of $1.505 \pm 0.005$~d, very close to the orbital period, and several lower peaks at periods between $\approx 1.2$ and $\approx 3$~d. Putting the residuals in phase with the orbital period, we see the occultation of the planet and a possible phase-dependence of the reflected light (see B12). \subsection{Longitude distribution of active regions and stellar differential rotation} \label{spot_model_res} \begin{figure*}[!t] \centerline{ \includegraphics[width=12cm,height=18cm,angle=90]{spot_contours2_occ_spots_final.ps}} \caption{Distribution of the spot-filling factor vs. longitude and time for $Q=1.6$. The values of the filling factor were normalized to their maximum $f_{\rm max} =0.01553$ with orange-yellow indicating the maximum and dark blue the minimum (see the colour scale on the right lower corner of the figure for the correspondence between the colour and the normalized filling factor). Note that the longitude scale is extended beyond $0^{\circ}$ and $360^{\circ}$ to help following the migration of the starspots. The tracks of the five spots occulted during the planetary transits after D11 are also reported. The open circles mark the time intervals of eight transits after which the same spots are detected again during the transits. The straight lines connecting the circles trace the migration of those spots in our reference frame; each line is labelled with the name of the corresponding spot, as indicated in Fig.~11 of D11. } \label{long_distr} \end{figure*} The distribution of the spot-filling factor $f$ vs. the longitude and the time is plotted in Fig.~\ref{long_distr}. The longitude zero corresponds to the point intercepted on the stellar photosphere by the line of sight to the centre of the star at BJD~2454964.5109, i.e., the sub-observer point at the initial epoch. The reference frame rotates with the star with a fixed period of $12.01$~d and the longitude increases in the same direction as the stellar rotation and the orbital motion of the planet. This is consistent with the reference frames adopted in our previous studies \citep[e.g., ][]{Lanzaetal09a,Lanzaetal09b}, but does not allow a direct comparison of the mapped active regions with the dips in the light curve. Our map shows that the individual starspots evolve with timescales of tens of days, which makes it difficult to trace the evolution and migration of the active regions in an unambiguous way. Nevertheless, two main active longitudes, where individual spots form and evolve, can be identified with confidence and appear to rotate on the whole with the rotation period of our reference frame. The migration rate of individual spots or groups of spots within each longitude is variable, which suggests that individual spots are forming at different latitudes on a differentially rotating star or, alternatively, there are several spots at close longitudes that are evolving to mimic spot migration. Unfortunately, there is no information on the starspot latitude from our mapping technique and even the hemisphere cannot be determined because the inclination is close to $90^{\circ}$ (cf. Sect.~\ref{spotmodel}). Nevertheless, some additional information can be extracted in this particular case from the occultation of the spots by the planet during transits. Specifically, we can exploit the results of D11 to constrain the latitude of some of the starspot trails that are seen in Fig.~\ref{long_distr}. \citet{Desertetal11} identified five starspots that are repeatedly occulted every eight transits and labelled them A, B, C, D, and E (cf. their Figs. 11 and 12). The transit profile distortions associated with spot C are barely visible in their Fig.~11, thus spot C is certainly significantly smaller than the other four spots. We plot in Fig.~\ref{long_distr} the migration of the starspots detected by D11. The initial epoch $E(0)$ of D11 is equal to the epoch of the first mid transit as reported in their Table~3 (D\'esert, priv. comm.) so we can easily convert their phases of maximum starspot visibility into longitudes in our reference frame. We find a good association between their starspots A, B, D, and E and some trails of spots found with our approach. For starspot C we find almost no coincidence, as expected owing to its smaller spotted area. The resolution of the transit-mapping method is of the order of some degrees \citep[cf., e.g., ][ and references therein]{Silvavalioetal10,SilvavalioLanza11} while that of our mapping based on the out-of-transit light curve is approximately $40^{\circ}-50^{\circ}$. Therefore, we are not able to resolve the occulted spots with the same detail as D11, especially when there are several small spots close in longitude. Note that a detailed comparison of the spot maps based on the in-transit and out-of-transit photometry is outside the scope of the present work because it requires a detailed modelling of the transit profile distortions induced by occulted starspots \citep[see ][ for such a detailed comparison for CoRoT-2]{SilvavalioLanza11}. Here we limit ourselves to exploiting the preliminary modelling of D11 to find that our trails of spots with a positive migration rate, i. e., rotating faster than our reference frame, are located inside the belt that is occulted by the planet. A faster migration is particularly evident in the case of our spot trails associated with their spots B and D, at least for a significant part of the considered time interval. On the other hand, the intrinsic evolution of the pattern makes the determination of the migration rates of our spots associated with their spots A and E less certain, although a positive migration is suggested. On the whole, these results indicate that the occulted spots located at low latitudes are rotating faster than the overall pattern of the active longitudes, thus giving evidence of a solar-like differential rotation in Kepler-17, i.e., with the equator rotating faster than the poles. Finally, we note that the time resolution of our spot models is not adequate to look for a possible modulation of the stellar activity with the orbital period of its close-in massive planet, as suggested by \citet{Shkolniketal08} or \citet{Lanza11}. Given the short orbital period of the planet, a different approach should be used to search for signatures of a possible star-planet interaction using short-cadence data, as for CoRoT-2 \citep[cf., e.g., ][]{Paganoetal09}. \subsection{Variation of the spotted area} The lifetimes of the active longitudes traced in Fig.~\ref{long_distr} range from $\approx 75$ to $\approx 330$ days and possibly longer, given the limited time interval covered by our observations. Individual active regions have a lifetime of about a few tens of days. This is similar to what we observe in the Sun where complexes of activity consisting of several active regions, forming approximately around the same longitude, have lifetimes up to $5-6$ months. The duration of the light anomalies observed during the transits reaches to $\approx 0.02$ in phase units, indicating a size of the largest occulted active regions of at least $\approx 40^{\circ}$ in longitude, i.e., $3-4$ times that of the largest sunspot groups. The variation of the total starspot area vs. the time is plotted in Fig.~\ref{total_area}. The error bars have a semi-amplitude of three standard deviations as derived from the uncertainty of the photometry, but systematic errors associated with the assumptions of the data reduction and the model are not included. Specifically, the variations on time scales longer than the duration of one quarter, i.e., $90-100$ days, cannot be reconstructed from \emph{Kepler} photometry because of the flux jumps from one quarter to the next (see Sect.~\ref{observations}). Therefore, we are limited to study the variations on timescales shorter than three months. Moreover, the presence of gaps in the observations can introduce systematic errors in the measurement of the spotted area. This is the case of the gap beginning at BJD~2454997.982 for a duration of 5.0268~d. Since the time interval adopted for our individual best fits is $\Delta t_{\rm f}= 8.733$~d, this loss of data implies a systematically lower value of the total spotted area because the ME regularization removes spots at the longitudes not constrained by the observations. Fortunately, the other gaps in the photometric time series are much shorter than $\Delta t_{\rm f}$, thus no other value of the area is significantly affected. To show the distribution of the gaps, we plot line segments with a length equal to their duration at the level 0.04 in Fig.~\ref{total_area} considering all interruptions with a duration longer than 24 hours. In the analysis presented below, we discarded the area value at BJD 2455003.81 because it is affected by the gap that started at 245997.982. The Lomb-Scargle periodogram of the entire time series of the area values is plotted in Fig.~\ref{pow_total_area} (solid line) together with the power level corresponding to a false-alarm probability of 0.01 as evaluated according to \citet{HorneBaliunas86}. The main peak corresponds to a period of $47.1 \pm 4.5$~d and has a false-alarm probability (FAP) of $2.0$ percent. The value of the FAP was confirmed by performing an analysis of 50,000 random Gaussian noise time series with the same sampling as our area data series. We also plot the periodogram of the time interval from BJD~2455230.869 to 2455457.929 (dashed line) because we see a regular oscillation of the total spotted area with a period of $\approx 48$~d during that interval in Fig.~\ref{total_area}. The main periodogram peak corresponds to a period of $48.2 \pm 9.0$~d with an FAP of $0.44$ percent. The time variation of the frequency of the spotted area modulation is best represented by means of a wavelet amplitude. We plot in Fig.~\ref{wavelet_total_area} the amplitude of the Morlet wavelet versus the period and the time \citep[see, e.g., ][ for details]{HempelmannDonahue97}. The wavelet parameters are adjusted to give a time resolution of $\approx 100$~d for a period of about $48$~d and a relative period resolution of $\Delta P / P \approx 0.06$. We see that in the initial part of the dataset there is a periodicity of $\approx 30$~d that corresponds to secondary peaks in the periodogram of the whole time series whose false-alarm probability is $ > 50$ percent (cf. Fig.~\ref{pow_total_area}, where the frequency resolution is better than in the case of the wavelet). On the other hand, during the second half of the time series we see a clear periodicity of $\approx 50$~d, which corresponds to the significant peak in the periodogram. We conclude that the total spotted area of Kepler-17 showed an oscillation with a period of $ 47.1 \pm 4.5$~d. This behaviour is reminiscent of the short-term oscillations of the total sunspot area found close to the maxima of some of the 11-year solar cycles. They were called Rieger cycles because they were first detected in the periodicity of occurrence of large solar flares by \citet{Riegeretal84}. In the Sun, the periodicity is $\approx 160$~d with small variations from one cycle to the other, although only some of the sunspot maxima show evidence of this short-term periodicity \citep{Oliveretal98,KrivovaSolanki02,Zaqarashvilietal10}. A behaviour similar to that of Kepler-17 was found in CoRoT-2, a G7V star that showed a Rieger-type cycle in the variation of its spotted area with a period of $28.9 \pm 4.3$~d \citep{Lanzaetal09a}. \begin{figure}[t] \centerline{ \includegraphics[width=6cm,height=9cm,angle=90]{spot_area_1.ps}} \caption{Total spotted area as derived from the regularized ME models vs. time for $Q=1.6$. The lower horizontal ticks mark the gaps in the photometric time series longer than 24 hours.} \label{total_area} \end{figure} \begin{figure}[t] \centerline{ \includegraphics[width=6cm,height=8cm,angle=90]{area_periodogram.ps}} \caption{Lomb-Scargle periodogram of the variation of the spotted area. The solid line gives the normalized power vs. the period for the whole time interval, while the dashed line gives the power for the time interval from BJD 2455230.8696 to BJD 2455457.9294 with the same normalization as adopted for the periodogram of the whole interval. The horizontal dotted line marks the 99 percent confidence level (FAP = 0.01). } \label{pow_total_area} \end{figure} \begin{figure}[b] \centerline{ \includegraphics[width=6cm,height=8cm,angle=90]{wavelet_area_var_final.ps}} \caption{Amplitude of the Morlet wavelet of the total spotted area variation vs. the period and the time. The amplitude was normalized to its maximum value. Different colours indicate different relative amplitudes from the maximum (yellow) to the minimum (dark blue) as indicated in the colour scale in the right lower corner. } \label{wavelet_total_area} \end{figure} \section{Discussion and conclusions} \label{conclusions} The application of a spot model to reproduce the optical light curve of Kepler-17 shows that the spot pattern is almost stable for a timescale of $\approx 9$~d because the residuals to our best fits have a standard deviation of $\sigma \approx 3.3 \times 10^{-4}$ in relative flux units that is only $\approx 50$ percent greater than the mean error attributed to the photometric measurements by the \emph{Kepler} pipeline. Our facular-to-spotted area ratio $Q=1.6$ is significantly lower than the value $Q_{\odot}=9.0$ adopted for the modelling of the light curves of the Sun-as-a-star by \citet{Lanzaetal07}. However, this lower value of $Q$ is typical of sun-like stars that are more active than the Sun \citep[cf. ][]{Lanzaetal09a,Lanzaetal10,Lanzaetal11a,Lanzaetal11b} and suggests an increasing weight of the dark spots in the photometric variations as a star becomes more active, as indicated by the results of, e.g., \citet{Radicketal98} and \citet{Lockwoodetal07}. Our models found several active longitudes that clustered on opposite hemispheres with a separation of $\approx 180^{\circ}$ for more than half the observation interval. This explains the two peaks found in the periodogram of the light curve by D11, one corresponding to the rotation period and the other to its first harmonic. \citet{Bonomoetal12} suggested an age younger than $1.8$~Gyr, while D11 derived an age of $3.0 \pm 1.6$~Gyr for Kepler-17. As noted by B12, the age determined by means of standard gyrochronology is only $0.9 \pm 0.2$~Gyr, while considering the effects of the close-in planet on the evolution of the stellar angular momentum \citep{Lanza10}, an age of $1.7 \pm 0.3$~Gyr is estimated that seems more compatible with the age found from isochrone fitting. Since the rotation period of the star is longer than the orbital period, tides remove angular momentum from the orbit to spin up the star and lead to orbital decay. The timescale for the engulfment of the planet can be estimated according to \citet{OgilvieLin07} as $\tau_{a} \simeq 0.048 (Q^{\prime}_{\rm s}/10^{6})$~Gyr, where $Q^{\prime}_{\rm s}$ is the modified tidal quality factor of the star. This timescale is much shorter than the lifetime of the system on the main sequence if we adopt $Q^{\prime}_{\rm s} \approx 10^{6}$, i.e., the value derived from the observed circularization periods of close binary systems in open clusters of different ages. Together with the observations of several other stars with massive planets on very tight orbits, this suggests that $Q^{\prime}_{\rm s}$ should be much higher (i.e., the tidal dissipation much lower) in those star-planet systems than in close binary systems that consist of two main-sequence stars. The difference in the $Q^{\prime}_{\rm s}$ value could arise because the stellar rotation is far from being synchronized with the orbital motion of its planet. In this case, considering the dissipation of the tides inside the convection zone, \citet{OgilvieLin07} predicted $Q^{\prime}_{\rm s} \ga 5 \times 10^{9}$, which implies an infall timescale longer than the system lifetime. A lower limit on the value of $Q^{\prime}_{\rm s}$ can be set by an accurate timing of the transits over a time interval of a few decades because for $Q^{\prime}_{\rm s}= 10^{6}$ we expect a variation of the orbital period of $\Delta P_{\rm orb}/P_{\rm orb} \approx 5 \times 10^{-8}$ in ten years. It produces a variation of $\approx 8$~s in the epoch of the mid transit in 10 years. The accuracy reported by D11 is $\pm \, 2.4$~s for their initial transit epoch, implying that a $Q_{\rm s}^{\prime}$ of about $10^{6}$ should give an orbital period acceleration detectable in a few years with a space-borne photometer. The recently approved extension of the \emph{Kepler} mission till 2016 is therefore an interesting opportunity to perform such measurements. A model of the light perturbations that are caused by the spots occulted during the transits may possibly be used to improve the accuracy since D11 found small $O-C$ timing oscillations with a period of approximately half the stellar rotation period, i.e., likely associated with starspots on opposite stellar hemispheres (cf. their Fig.~10). We can estimate an approximate lower limit to the differential rotation of Kepler-17 finding $\Delta \Omega/\Omega \approx 0.10-0.16$ from the migration rates of the different trails of spots as seen in Fig.~\ref{long_distr}. Given the rapid evolution of the individual spots, this value is not only approximate, but depends critically on the way individual spot trails are identified. Therefore, we caution to regard this estimate as remarkably uncertain (cf. Sect.~\ref{spot_model_res}). For comparison, in k$^{1}$~Ceti, a dwarf star with a spectral type G5V and a mean rotation period of $\approx 9$~d, \citet{Walkeretal07} estimated a relative differential rotation amplitude of $\approx 0.090 \pm 0.006$ by modelling the optical photometry obtained by the satellite MOST. A result similar to that of Kepler-17 was found by \citet{Crolletal06} for $\epsilon$~Eridani, a K2V star with a rotation period of $\approx 11.4$~d, showing a relative amplitude of the differential rotation of $0.11 \pm 0.03$. On the other hand, the more rapidly rotating star CoRoT-2 ($P_{\rm rot} \approx 4.52$~d) shows very little surface differential rotation with a relative amplitude smaller than $\approx 1$ percent \citep{Lanzaetal09a}. In the young ($\approx 50$~Myr) K2V star KIC~8429280, \citet{Frascaetal11} found $\Delta \Omega/ \Omega \approx 0.05$ while for the G2V stars KIC~7985370 and KIC~7765135, which have an age of $100-200$~Myr, \citet{Frohlichetal12} found $\Delta \Omega / \Omega \approx 0.07-0.08$. The differences can be attributed to the dependence of the differential rotation amplitude on the stellar rotation period and effective temperature as well as to the limited range of latitudes covered by starspots that can vary along the activity cycle. Moreover, the amplitude of the differential rotation could change along the activity cycle in rapidly rotating, highly active stars, such as the above mentioned \emph{Kepler} targets that have $P_{\rm rot}$ ranging from 1.2 and 2.8~d \citep[cf. ][]{Frascaetal11,Frohlichetal12}. Therefore, individual results coming from spot modelling can be compared only in a statistical sense. Considering the general trend found by \citet{Barnesetal05}, the amplitude of the differential rotation estimated for Kepler-17 appears within the expected range for a star with its rotation period and effective temperature. Nevertheless, a more extended series of data is required to derive a firm conclusion on this point as well as on a possible influence of the planet on spot activity as conjectured by, e.g., \citet{Lanza08,Lanza11}. An interesting result of our analysis is the short-term spot cycle with a period of $47.1 \pm 4.5$~d, which is clearly detected in the second half of the data series. This phenomenon is reminiscent of the solar Rieger cycles because of its timescale and transient nature. The periodicity of the spotted area variations is close to $4$ times the mean rotation period of the star, while in the Sun it is $\approx 6$ times and in CoRoT-2 is $\approx 6.5$ times the rotation period. Those oscillations of the spotted area may be associated with hydromagnetic Rossby-type waves propagating in the upper part of the convection zone or at the interface between the radiative interior and the convection zone where the dynamo is probably located \citep{Lou00,Zaqarashvilietal10}. Since only a few examples of stars displaying Rieger-type cycles are known \citep[cf., ][]{Massietal05,Lanzaetal09a}, the new results on Kepler-17 are particularly interesting for a better understanding of this phenomenon in the framework of the solar-stellar connection. As new \emph{Kepler} data become available, it will be possible to refine the conclusions of the present study by investigating a longer time span. This is needed to assess the duration and the frequency of the Rieger-type oscillations in the total starspot area as well as to derive a definite conclusion on the possible impact of the planet on stellar activity. \begin{acknowledgements} The authors wish to thank an anonymous referee for several valuable comments, the NASA and the \emph{Kepler} team for giving public access to the \emph{Kepler} data. A.~S.~Bonomo gratefully acknowledges support through an INAF/HARPS-N fellowship. Active star research and exoplanetary studies at INAF-Osservatorio Astrofisico di Catania and Dipartimento di Fisica e Astronomia dell'Universit\`a degli Studi di Catania are funded by MIUR ({\it Ministero dell'Istruzione, dell'Universit\`a e della Ricerca}) and by {\it Regione Siciliana}, whose financial support is gratefully acknowledged. This research has made use of the ADS-CDS databases, operated at the CDS, Strasbourg, France. \end{acknowledgements}
\section{Introduction} \label{sec:intro} Computational topology lies properly within the broad scope of applied general topology and depends upon a novel integration of the `pure mathematics' of topology with the `applied mathematics' of numerical analysis. Computational topology also blends general topology, geometric topology and knot theory with computer visualization and graphics, as presented here. For the computational cases considered here, ideas from general topology, geometric topology and knot theory are complemented by numeric arguments in a novel integration of the `pure' field of topology with the `applied' focus of numerical analysis. Additionally, aspects of computer visualization and graphics are incorporated into the proofs. Much of this work was motivated by modeling biological molecules, such as proteins, as knots for visualization synchronized to simulations of the writhing of these molecules. Attention is restricted to when $\mathcal{C}$ is closed, implying that $\mathcal{L}$ is also closed. As $\mathcal{C}$ is created from $\mathcal{L}$, it is natural to ask which topological characteristics are shared by these two curves, particularly as the control polygon often serves as an approximation to the B\'ezier curve in many practical applications. However, topological differences between a B\'ezier curve and its control polygon can exist and it is natural to develop counterexamples to show these topological differences. These counterexamples extend beyond related results, while we introduce new computational methods to generate additional counterexamples. In Section~\ref{sec:ex1}, we present a counterexample of B\'ezier curve and its control polygon being homeomorphic, yet not ambient isotopic. To develop this counterexample, we created and used a computer visualization tool to study topological relationships between a B\'ezier curve and its control polygon. We viewed the images to motivate formal proofs, which also rely on numerical analysis and geometric arguments. In Section~\ref{sec:eqsimnot}, we present numerical techniques to create a class of topological counterexamples -- where a B\'ezier curve and its control polygon are not even homeomorphic, as the B\'ezier curve is self-intersecting while the control polygon is simple. We exhibit self-intersection by a numerical method, which finds the roots of a system of equation. We freely admit that these roots are not determined with infinite precision, but such calculations on polynomials of degree 6, as in these examples, typically elude precise calculation. We argue that two primary values of our method for these approximated solutions are \begin{enumerate} \item as a catalyst to alternative examples that may admit infinite precision calculation and rigorous topological proofs, and \item as having specified digits of accuracy -- typically crucial for acceptable approximations in computational mathematics and computer graphics. \end{enumerate} Using the visualization tool described in Section~\ref{sec:vizb}, we viewed many examples where the B\'ezier curve was simple, while its control polygon was equilateral and simple. It is well known that B\'ezier curve can be self-intersecting even when its control polygon is simple, but we conjectured that the added equilateral hypothesis would imply that both curves were simple. While this visual evidence was suggestive, we present a general numerical approach in Section~\ref{sec:eqsimnot} that supports a contrary interpretation. This example provides guidance for designing appropriate approximation algorithms for computer graphics. \subsection{Mathematical Definitions} \label{ssec:mdef} The definitions presented are restricted to $\mathbb{R}^3$, as sufficient for the purposes of this paper, but the interested reader can find appropriate generalizations in published literature. \begin{defn} \label{def:knot} A knot \cite{Armstrong1983} is a curve in $\mathbb{R}^3$ which is homeomorphic to a circle. \end{defn} Knots are often described by a {\em knot diagram} \cite{Livingston1993}, which is a planar projection of a knot. Self-intersections in the knot diagram correspond to {\em crossings} in the knot, where each crossing has one arc that is an {\em undercrossing} and an {\em overcrossing}, relative to the direction of projection. Homeomorphism is an equivalence relation over point sets and does not distinguish between different embeddings, while ambient isotopy is a stronger equivalence relation which is fundamental for classification of knots. \begin{defn} Let $X$ and $Y$ be two subspaces of $\mathbb{R}^3$. A continuous function \[ H:\mathbb{R}^3 \times [0,1] \to \mathbb{R}^3 \] is an {\bf ambient isotopy} between $X$ and $Y$ if $H$ satisfies the following: \begin{enumerate} \item $H(\cdot, 0)$ is the identity, \item $H(X,1) = Y$, and \item $\forall t \in [0,1], H(\cdot,t)$ is a homeomorphism from $\mathbb{R}^3$ onto $\mathbb{R}^3$. \end{enumerate} The sets $X$ and $Y$ are then said to be {\bf ambient isotopic}. \label{def:aiso} \end{defn} \begin{defn}\label{def:c} Denote $\mathcal{C}(t)$ as the parameterized B\'ezier curve of degree $n$ with control points $P_m \in \mathbb{R}^3$, defined by $$ \mathcal{C}(t)=\sum_{i=0}^{n}{B_{i,n}(t)P_i}, t\in[0,1] $$ where $B_{i,n}(t) = \left(\!\!\! \begin{array}{c} n \\ i \end{array} \!\!\!\right)t^i(1-t)^{n-i}$. \end{defn} \subsection{Related Literature} \label{ssec:relwork} A B\'ezier curve and its control polygon may have substantial topological differences. It is well known that a B\'ezier curve and its control polygon are not necessarily homeomorphic \cite{Piegl}. Recently, it was shown that there exists an unknotted B\'ezier curve with a knotted control polygon \cite{Bisceglio}. A specific dual example has also been shown \cite{Carlo} of a knotted B\'ezier curve with an unknotted control polygon. However, the methodology was a visual construction without formal proof and is not easily generalized. Software, {\em Knot\_Spline\_Vis}, developed by authors Marsh and Peters, was used to find another example, where the methodology can more easily be generalized and {\em Knot\_Spline\_Vis} is publicly available \cite{tjp-web}. Much of the motivation for considering these counterexamples came from applications in scientific visualization \cite{JKSP10,TCS08,JMPR11}. A primary focus was to establish tubular neighborhoods of knotted curves so that piecewise linear (PL) approximations of those curves within those neighborhoods remained ambient isotopic to the original curves. This was initially considered for approximations used in producing {\em static} images of these curves, but it became readily apparent that these same neighborhoods also provided bounds within which many perturbations of these models remained ambient isotopic under these movements. That theory is being applied to dynamic visualizations of molecular simulations, where the neighborhood boundaries permit convenient warnings as to an impending self-intersection, as of possible interest to biologists and chemists who are running the simulations. Previous work \cite{KnotPlot} in knot visualization provides a rich set of data for PL knots, where each edge is of the same length. The interface of {\em Knot\_Spline\_Vis} was designed to import this equilateral PL knot data and then generate the associated B\'ezier curves. This matured into an empirical study of dozens of cases, where all examples examined yielded simple B\'ezier curves for these simple equilateral control polygons. This raised the question of whether the presence of equilateral edges in the control polygon might be a sufficient additional hypothesis to ensure homeomorphic equivalence with the B\'ezier curve, as the previously cited examples \cite{Piegl} did not have equilateral control polygons. \section{Unknotted $\mathcal{L}$ with knotted $\mathcal{C}$} \label{sec:ex1} In order to produce a knotted B\'ezier curve with an unknotted control polygon, we invoked {\em Knot\_Spline\_Vis} with an example (Figure~\ref{fig:pre-unknotpoly}) of an unknotted B\'ezier curve, where the total curvature appeared to be larger than $4\pi$. (The total curvature being larger than $4\pi$ is a necessary condition of knottedness.) \begin{figure}[h!] \centering \subfigure[Unknotted $\mathcal{C}$ \& The $\mathcal{P}$] { \includegraphics[height=5cm]{unknotcandp} \label{fig:unkCP} } \subfigure[Zoomed-in View of $\mathcal{C}$] { \includegraphics[height=5cm]{unknotci} \label{fig:unkview} } \caption{Visual experiments} \label{fig:pre-unknotpoly} \end{figure} We experimented on this example by moving control points to construct a B\'ezier curve that visually appeared to be knotted. We then moved control points to unknot the control polygon while keeping the B\'ezier curve knotted. In the end we obtained a B\'ezier curve and the control polygon (of degree $10$) (Figure~\ref{fig:k}) defined by the control points $\{P_0,P_1,\cdots,P_9,P_0\}$ listed below: {\footnotesize $$(-5.9,4.7,-6.2), (10.3,-1.1,8.9), (-2.6,-12.4,-6.3), (-10,7,-0.3),$$ $$ (1.9,-12,-0.6), (11.2,7.5,-7.6), (-15.3,-1.7,-4.1), (-11.7,20,3.5), $$ $$(17.9,-1.1,2.9), (2.9,-13.7,4.8), (-5.9,4.7,-6.2).$$} The 3D visualization offers only suggestive evidence that the above B\'ezier curve is knotted while the control polygon is unknotted. We provide rigorous mathematical proofs of these properties in Sections~\ref{ssec:p1a} and \ref{ssec:p1b}. Generally, proving knottedness or unknottedness can be difficult \cite{Hass1999}, but is accessible for the counterexample here. \subsection{Proofs of the B\'ezier curve being knotted} \label{ssec:p1a} We prove that the B\'ezier curve is a trefoil\footnote{A trefoil is a knot with three alternating crossings\cite{Livingston1993}.}. We orthogonally project the curve onto $x$-$y$ plane. We then shown that there are three self-intersections in the projection and these intersections are alternating crossings in 3D. Since projections preserve self-intersections, this curve can have no more than 3 self-intersections, but these self-intersections in $x$-$y$ plane are shown to have pre-images that are 3D crossings, so the original curve has no self-intersections. \begin{figure}[h!] \begin{center} \includegraphics[height=7cm]{unknotpoly-dprojection} \end{center} \caption{The 2D projection of the knot}\label{fig:k-2d} \end{figure} Since the 2D curve in Figure~\ref{fig:k-2d} has degree $10$, we use a numerical method implemented by MATLAB function `fminsearch' to find the parameters where the curve intersects with itself. We provide the numerical codes in Appendix~\ref{app:selfint}, and we provide the data used to find these parameters in Appendix~\ref{data:selfint}. The pairs of parameters (labeled in order as $t_1,\ldots,t_6$) of the self-intersections are listed below: $$[t_1=0.0306,t_4=0.5573], [t_2=0.1573,t_5=0.9244], [t_3=0.3731,t_6=0.9493].$$ Next we prove that these 2D intersections are projections from three alternating crossings in 3D. The above parameters are substituted into the B\'ezier curve (Definition~\ref{def:c}) to get pairs of points (numerical codes for this calculation are in Appendix~\ref{app:cr}): {\small $$[\mathcal{C}(t_1)=(-2.0539,2.8001,-2.6929), \mathcal{C}(t_4)=(-2.0530,2.7987,-2.0143)],$$ $$[\mathcal{C}(t_2)=(0.4376,-2.5212,-0.0576),\mathcal{C}(t_5)=(-0.4364,-2.5206,-0.5547)],$$ $$[\mathcal{C}(t_3)=(-1.3613,-1.4239,-2.2944),\mathcal{C}(t_6)=(-1.3624,-1.4232,-1.9067)].$$} The alternating crossings follow from comparing the z-coordinates in each pair. Precisely, according to the parameters given above, the tracing of the six points in order is $$\mathcal{C}(t_1), \mathcal{C}(t_2), \mathcal{C}(t_3), \mathcal{C}(t_4), \mathcal{C}(t_5), \mathcal{C}(t_6),$$ and the crossings at these points are $$under, over, under, over, under, over.$$ \subsection{Proof of the control polygon being unknotted} \label{ssec:p1b} To prove that the control polygon $P=(P_0,P_1,\cdots,P_9,P_0)$ is an unknot, it is necessary to show that $P$ is simple and unknotted. We directly tested each pair of segments of $P$ for non-self-intersection and those calculations can be repeated by the interested reader. We prove unknottedness using a 3D {\em push}, as the obvious generalization of a 2D function from low-dimensional geometric topology \cite{Bing1983}. We restrict attention to a {\em median push}, as defined below. The full sequence of 5 median pushes is explicated, where the first 4 median pushes are equivalently described by Reidemeister moves \cite{Livingston1993}. \begin{figure}[h!] \centering \subfigure[The initial control polygon] { \includegraphics[height=4cm]{unknot0-push} \label{fig:unknot0} } \subfigure[After the 1st push] { \includegraphics[height=4cm]{unknot1-push} \label{fig:unknot1} } \subfigure[After the 2nd push] { \includegraphics[height=4cm]{unknot2-push} \label{fig:unknot2} } \subfigure[After the 3rd push] { \includegraphics[height=4cm]{unknot3-push} \label{fig:unknot3} } \subfigure[After the 4th push] { \includegraphics[height=4cm]{unknot4-push} \label{fig:unknot4} } \subfigure[The unknot after these pushes] { \includegraphics[height=4cm]{unknot5} \label{fig:unknot5} } \caption{Pushes and PL curves in 3D} \label{fig:pushes} \end{figure} \begin{defn} \label{def:mp} Suppose $\triangle{P_{j-1}P_jP_{j+1}}$ is a triangle determined by non-collinear vertices $P_{j-1}, P_j$ and $P_{j+1}$ of $P$. Push the vertex $P_j$, along the corresponding median of the triangle to the middle point of the side $P_{j-1}P_{j+1}$. We call this function a median push.\end{defn} We depict the sequence of pushes used in Figure~\ref{fig:pushes}, showing the 3D graphs of the PL space curves after the median pushes. These graphs show, at each step, which vertex is pushed and its image. For example, in Figure~\ref{fig:unknot0}, the vertex $P_3$ is pushed to $P'_3$ and the resultant polygon after this push is shown by Figure~\ref{fig:unknot1}. Figures~~\ref{fig:unknot0}~\ref{fig:unknot1}~\ref{fig:unknot2} and~\ref{fig:unknot3} have corresponding Reidemeister moves, as those pushes eliminate at least one crossing. Using the published notation \cite{Livingston1993}, Figure~\ref{fig:unknot0} depicts a Reidemeister move of Type 2b. Similarly, Figure~\ref{fig:unknot1} depicts a Reidemeister move of Type 1b; Figure~\ref{fig:unknot2} has Type 2b and Figure~\ref{fig:unknot3} has a move of Type 2b, followed by a move of Type 1b. The final push to achieve Figure~\ref{fig:unknot5} does not correspond to any Reidemeister move as no crossings are changed, but it is included to have a polygon with only five edges, which necessarily must be the unknot \cite{adams2004knot}. We prove that $P$ is unknotted by showing that $P$ is ambient isotopic to the unknotted PL curve shown in Figure~\ref{fig:unknot5}. As a sufficient condition, we show that the pushes do not cause intersections \cite{J.W.Alexander_G.B.Briggs1926}. (We gained significant intuition for specifying the sequence of pushes by visual verification with our 3D graphics software capabilities to translate, rotate and zoom the images.) We now present the formal arguments. Consider Figure~\ref{fig:unknot0}, where $P_3$ is pushed to $P'_3$ (the middle point\footnote{It is not necessary to push $P_3$ to the middle point. Any point along $\overrightarrow{P_2P_4}$ would suffice.} of $\overrightarrow{P_2P_4}$). We show that any segments other than $\overrightarrow{P_2P_3}$ and $\overrightarrow{P_3P_4}$ of $P$ do not intersect the triangle $\triangle{P_2P_3P_4}$ or the triangle interior. We parameterize each segment by: $$\overrightarrow{P_iP_{i+1}}: P_i+(P_{i+1}-P_i)t, \ t\in[0,1]$$ for $i=0,1,\cdots,9$ and let $P_{10}=P_0$. Then the points given by $$P_3+a(P_2-P_3)+b(P_4-P_3),$$ for $a,b \geq 0$ and $a+b \leq 1$ are on the $\triangle{P_2P_3P_4}$ and contained in its interior. Hence $P_i+(P_{i+1}-P_i)t$ intersects $\triangle{P_2P_3P_4}$ or its interior if and only if \begin{align}\label{eq:triangle}P_i+(P_{i+1}-P_i)t=P_3+a(P_2-P_3)+b(P_4-P_3)\end{align} has a solution for some $t\in [0,1]$ and $a,b \geq 0$ and $a+b \leq 1$. For each $i=0,1,\cdots,9$, we solve Equation~\ref{eq:triangle} (a system of $3$ linear equations) for $a,b$ and $t$ with the above constraints. Calculations show there is no solution for each system, and hence $P_i+(P_{i+1}-P_i)t$ does not intersect $\triangle{P_2P_3P_4}$ or its interior for each $i=0,1,\cdots,9$. Thus it follows that the push does not cause any intersections. Similar computations verify that the other pushes do not cause any intersections, thus establishing the ambient isotopy. \section{Equilateral, simple $\mathcal{L}$ with self-intersecting $\mathcal{C}$} \label{sec:eqsimnot} We visualized many cases of simple, closed equilateral control polygons\footnote{Throughout this section, all control polygons are simple.} that all appeared to have unknotted B\'ezier curves. Many of these control polygons were nontrivial knots. Prompted by this visual evidence, we conjectured that any simple, closed equilateral control produces a simple B\'ezier curve, where some examples are shown in Figure~\ref{fig:examples-equilateral}. We now present numerical evidence to the contrary. As noted in Section~\ref{sec:intro}, the degree 6 polynomials make precise computation difficult, so we do not provide a completely formal proof, independent of numerical methods. A legitimate concern is whether the numerical approximation produces coordinates for a `near' intersection, subject to the accuracy of the floating point computation. We cannot refute that possibility, but we argue that in the context of graphical images, the level of approximation produced is often sufficient. In particular, we have parameters in the code to adjust the number of digits of accuracy. This is level of user-defined precision is often accepted as sufficient for visualization \cite{JKSP10}. The user can then set the graphical resolution so that points that are determined to be the same within some acceptable numerical tolerance will also appear within the same pixel. \begin{figure}[h!] \begin{center}\includegraphics[height=7cm]{examples-equilateral}\end{center} \caption{Unknotted B\'ezier curves with equilateral control polygons}\label{fig:examples-equilateral} \end{figure} \subsection{Intuitive overview} \label{ssec:iover} We create many examples to test and retain only those that satisfy all three criteria listed in Section~\ref{ssec:num}. We begin the creation of a closed equilateral polygon by setting $P_0=(0,0,0)$. We then take $\{q_0,q_1,\cdots,q_{n-1}\}$ (Equation~\ref{eq:q}) from the unit sphere so that $$P_1=P_0+q_0,\ \ \ P_2=P_1+q_1,\ \ \ \cdots,\ \ \ P_n=P_{n-1}+q_{n-1}.$$ We ensure that the polygon is closed in Section~\ref{ssec:num} item $(3)$. We consider a sufficient and necessary condition for a B\'ezier curve being self-intersecting (Equation~\ref{eq:S2}). Since we want the equilateral polygon to define a B\'ezier curve that is self-intersecting, we not only select $\{q_0,q_1,\cdots,q_{n-1}\}$ from the unit sphere as above, but also select them such that Equation~\ref{eq:S2} is zero for some parameters $s$ and $t$. Consequently the set of control points generated determines a closed equilateral control polygon and a self-intersecting B\'ezier curve. \subsection{Necessary and sufficient condition for self-intersection} We rely upon the following published equation \cite{Andersson1998} for necessary and sufficient conditions for self-intersection of a B\'ezier curve \begin{equation} S(s,t)=\frac{1}{n}\frac{\mathcal{C}(1-s)-\mathcal{C}(t)}{(1-s)-t}, \label{eq:Bselfint} \end{equation} with the domain $D=\{(s,t): s+t < 1, s,t \geq 0, (s,t) \neq (0,0)\}$. A B\'ezier curve defined by $\mathcal{C}(t)$ is self-intersecting if and only if there exist $s$ and $t$ in the domain $D$ such that $S(s,t)=0$, , with an alternative formulation\footnote{One needs to write out the \cite[Equation (6)]{Andersson1998} to obtain Equation~\ref{eq:S2}.} given by \begin{align}\label{eq:S2} S(s,t) = \frac{1}{n} \sum_{i=0}^{n-1} \sum_{j=0}^{n-1-i} \binom{n-1-i}{j} s^{n-1-i-j} (1-s)^j \sum_{k=0}^i \binom{i}{k}(1-t)^{i-k}t^kq_{j+k},\end{align} where \begin{align}\label{eq:q}q_i=P_{i+1}-P_{i}\ \ \ \ for\ i=\{0,1,\cdots,n-1\}.\end{align} \subsection{A representative counterexample generated} \label{sec:ex} We present a single numerical counterexample, noting that while only one counterexample is presented, the numerical algorithm implemented can be used to find many such examples. We list six distinct control points (the seventh control point is equal to the first control point) that determine an equilateral simple $\mathcal{L}$ and a self-intersecting $\mathcal{C}$ (shown in Firgure~\ref{fig:knots1}), as generated by the algorithm described in detail in Section~\ref{ssec:num}. $$(0,0,0), (0.0305,0.0810,0.9962), (-0.2074, -0.2671,1.9030), $$ $$ (-0.1792,-0.3402,0.9063), (0.0189, 0.0782,0.0185), (0.1557, 0.2329,-0.9600) .$$ We verify that $\mathcal{L}$ is simple by considering all pairwise intersections of the segments of this control polygon. The self-intersection of the B\'ezier curve occurs at a point that is numerically approximated as $$[s,t]=[0.2969,0.0633]$$ where correspondingly, $$S(s,t)=(-0.0003861,-0.000097,0.0001462) \approx (0,0,0).$$ The error occurs because of numerical round off on $s,t$ and the control points. \subsection{The numerical method for generating counterexamples} \label{ssec:num} We provide the numerical codes and data used in Appendix~\ref{app:eq}. Given a control polygon, we can determine whether a self-intersection of the B\'ezier curve occurs by determining whether $S$ (Equation~\ref{eq:S2}) has a root in $D$. We consider $S$ as a function $S(s,t,q)$ where $q=\{q_i\}_{i=0}^{n-1}$, so that finding a self-intersecting B\'ezier curve with an equilateral control polygon is equivalent to determining $s,t$ and $q$ such that the following are satisfied: \begin{enumerate} \item $S(s,t,q)=0$ where $s,t\in D$; \item $||q_i||=r$ for each $i \in \{0,1,\cdots,n-1\}$, where $||\cdot||$ is the Euclidean norm; \item $\sum_{i=0}^{n-1}{q_i}=\sum_{i=0}^{n-1}{P_{i+1}-P_i}=0$ since $P$ needs to be closed. \end{enumerate} We assume $r=1$ without loss of generality since the value of $r$ can be adjusted by scaling. Throughout the provided codes, $n$ is always the degree of the B\'ezier curve. We give the code for function $S$ in Appendix~\ref{app:S}, where $[s,t]$ is labeled as $u$ and $q$ as $[a,b,c]$. The function $SF$ (Appendix~\ref{app:SF}) takes parameters for $s,t$ and $q$ and outputs a floating point value. It is designed to be zero if and only if the above three conditions are satisfied simultaneously. Precisely since $q$ should be taken from the unit sphere, $SF$ assigns $a,b,c$ values given by $$a=sin(\phi)cos(\theta); $$ $$b=sin(\phi)sin(\theta);$$ $$c=cos(\phi),$$ where $\phi$ and $\theta$ represent input parameters. In order to satisfy Condition $(3)$ above, $q_{n-1}$ is set equal to $-\sum_{i=0}^{n-2}{q_i}$. But in this way, $||q_{n-1}||$ may fail to be equal to $1$. So we include the function $F$ (Appendix~\ref{app:SF}) to determine whether $||q_{n-1}||=1$. The function is designed to be $F=||q_{n-1}||-1$ such that $||q_{n-1}||=1$ if and only if $F=0$. Symbolically, \begin{align}\label{eq:SF} SF=||S||+||F||,\end{align} where $S$ is given by Equation~\ref{eq:S2} and $F=||q_{n-1}||-1$. Having the above three conditions satisfied simultaneously is equivalent to finding input values such that $SF=0$. Since $SF \geq 0$, the minimum of $SF$ is $0$. The function $SFminimizer$ (Appendix~\ref{app:SFminimizer}) uses $fminsearch$ (a numerical method integrated in MATLAB) to search for the minimum of $SF$, while returning this minimum and the corresponding values of $s,t$ and $q$. The user supplied initial guesses for $s,t$ and $q$ greatly influence the results, so we assign $SFminimizer$ randomized initial values $M$ times so that we get $M$ different minimums for different initial values. But no matter which initial values we use, as long as we can get ``a" minimum of $0$, then we get the equilateral control polygon which determines a self-intersecting B\'ezier curve. The data of finding the counterexample of Section~\ref{sec:ex} is included in Appendix~\ref{app:data}. \section{Visualizing B\'ezier curves \& their control polygons} \label{sec:vizb} To study the knot types of a control polygon and its B\'ezier curve, a knot visualization tool was developed, called {\em Knot\_Spline\_Vis}. The tool {\em Knot\_Spline\_Vis} takes a PL control polygon as input and displays both the PL curve and the associated B\'ezier curve. For these studies, the input was always restricted to be a PL curve of known knot type. The functions of {\em Knot\_Spline\_Vis} were designed to permit interactive studies of the topological relationships between a B\'ezier curve and its control polygon. The intent is to use these examples to stimulate mathematical conjectures as a prelude to formal proofs. Some of the standard graphical manipulation capabilities provided are illustrated in the following Figure~\ref{fig:rs} and~\ref{fig:zz}, where the rotation capabilities have been particularly useful to develop visual insights into the occurrence of self-intersections and crossings as fundamental for the study of knots. An editing window allows the user to change the coordinates of the control polygons, as shown in Figure~\ref{fig:move}. The software is freely available for download at the site \mbox{www.cse.uconn.edu/$\sim$tpeters}. \begin{figure}[h!] \centering \includegraphics[height=5cm]{ro} \caption{Rotating} \label{fig:rs} \end{figure} \begin{figure}[h!] \centering \includegraphics[height=5cm]{knot-10-z} \caption{Scaling} \label{fig:zz} \end{figure} \begin{figure}[h!] \centering \subfigure[Initial display] { \includegraphics[height=4cm]{ex941} } \subfigure[Some vertices moved] { \includegraphics[height=4cm]{ex941-moved} } \caption{Moving control points} \label{fig:move} \end{figure} \begin{figure}[h!] \centering \includegraphics[height=5cm]{sub} \caption{Subdivision} \label{fig:k10sub} \end{figure} The control polygon is an initial PL approximation to its associated B\'ezier curve. Subdivision algorithms \cite{Piegl} produce additional control points to further refine the original control polygon, converging in distance to the B\'ezier curve. Figure~\ref{fig:k10sub} illustrates performance of a standard subdivision technique, the de Casteljau algorithm \cite{G.Farin1990}. Users can specify a subdivision parameter. Figure~\ref{fig:k10sub} shows the case with parameter $\frac{1}{2}$. \section{Conclusions and Future Work} \label{sec:conc} We use numerical algorithms and graphical software to generate counterexamples regarding the embeddings in $\mathbb{R}^3$ for a B\'ezier curve and its control polygon. First, we present a class of counterexamples showing an unknotted control polygon for a knotted B\'ezier curve and provide complete formal proofs of this condition. We used a spline visualization tool, {\em Knot\_Spline\_Vis} for easy generation of many examples to gain intuitive understanding to formalize these proofs. Secondly, we used {\em Knot\_Spline\_Vis} in formulating the conjecture that any simple, closed equilateral control had a simple B\'ezier curve. We provide contrary numerical evidence regarding this conjecture, as a useful guide for many applications in computer graphics and scientific visualization. This work also shows the importance of continuing investigation \begin{itemize} \item theoretically, into an infinite precision proof for a counterexample of the simple, closed equilateral polygon conjecture, where these demands will likely require further, substantive mathematical innovation and \item practically, into a user interface for {\em Knot\_Spline\_Vis} to permit interactive 3D editing, as the current text driven interface was cumbersome. \end{itemize} \section{Web Posting of Supplemental Materials} \label{sec:web} Appendices listed below are posted on this webpage: \begin{center} \url{http://www.math.uconn.edu/~jili/Supplemental-materials.pdf} \end{center} \noindent \textbf{Appendix A:} Numerical codes for knottedness of $\mathcal{C}$ (Figure~\ref{fig:k}): \begin{itemize} \itemsep -2pt \item A.1 Codes for searching self-intersections in Figure~\ref{fig:k-2d}; \item A.2 Codes for determining under or over crossings; \item A.3 Data for searching intersections in Figure~\ref{fig:k-2d}. \end{itemize} \noindent \textbf{Appendix B:} Numerical codes for searching the Example of Figure~\ref{fig:eq}: \begin{itemize} \itemsep -2pt \item B.1 Codes for Equation~\ref{eq:S2}; \item B.2 Codes for Equation~\ref{eq:SF}; \item B.3 Codes for searching the roots of the system of Equations~\ref{eq:S2} and~\ref{eq:SF}; \item B.4 Data for finding the example shown in Figure~\ref{fig:eq}. \end{itemize} \pagebreak \bibliographystyle{plain}
\section{Introduction} Let $M$ be a complete Riemannian manifold of nonpositive sectional curvature. The \emph{geometric rank} of a geodesic $\gamma$ in $M$, denoted by $rk(\gamma)$, is the dimension of the vector space of parallel Jacobi fields along $\gamma$. Then the \emph{geometric rank} of $M$ is defined to be the minimum of $rk(\gamma)$ over all geodesics $\gamma$ in $M$. The celebrated rank-rigidity theorem, due to Ballmann (\cite{Bal}), and to Burns-Spatzier (\cite{BurSpa}), states that if $M$ has bounded nonpositive sectional curvature and finite volume, then the universal cover $\widetilde{M}$ is a flat Euclidean space, a symmetric space of non-compact type, a space of rank $1$ or a product of such spaces. In \cite{PraRag}, Prasad and Raghunathan introduced the notion of the algebraic rank, $\mathrm{rank}(G)$, of a group $G$. (See Section 2 for the definition.) Ballmann and Eberlein proved that if $\Gamma$ is the fundamental group of a complete Riemannian manifold $M$ of bounded nonpositive sectional curvature and of finite volume, then $\mathrm{rank}(\Gamma)$ is equal to the geometric rank of $M$ (see \cite{BalEbe}). By combining this with the rank-rigidity theorem, we have that, for a complete Riemannian manifold $M$ of bounded nonpositive sectional curvature and of finite volume, if $\widetilde{M}$ does not have an Euclidean factor and $\Gamma=\pi_{1}(M)$ has higher algebraic rank, either (1) $\Gamma$ is a lattice in a semi-simple Lie group of higher rank, (2) it has a finite index subgroup which splits as a direct product, i.e., $\Gamma$ is a virtually product, or (3) it acts on a product without being a virtual product. There is an analogous notion of geometric rank for $\mathrm{CAT}(0)$ spaces. A \emph{geometric flat of dimension $n$} in a complete $\mathrm{CAT}(0)$ space $X$ is a closed convex subset of $X$ which is isometric to the Euclidean $n$-space. A geodesic line $L$ is said to have \emph{rank one} if it does not bound a flat half-plane. A complete $\mathrm{CAT}(0)$ space $X$ is said to have \emph{higher geometric rank} if no geodesic in $X$ has rank one. Let $X$ be a complete $\mathrm{CAT}(0)$ space and $G$ be a group acting geometrically (i.e., properly and cocompactly by isometries) on $X$. In view of the Ballmann-Eberlein's result, it is natural to ask the similar question for $\mathrm{CAT}(0)$ spaces: \begin{conjecture} $G$ has higher algebraic rank if and only if $X$ has higher geometric rank. \end{conjecture} In this paper, we study the algebraic rank of various ${\mathrm{CAT}}(0)$ groups. They include right-angled Coxeter groups, right-angled Artin groups, relatively hyperbolic groups and groups acting geometrically on ${\mathrm{CAT}}(0)$ spaces with isolated flats. In Section 3, we prove that if $W$ is an infinite irreducible non-affine right-angled Coxeter group, then $\mathrm{rank}(W) =1$. As a corollary, we obtain a new proof for the question posed by M. Davis in \cite{Dav}, namely, $W$ cannot be commensurable to any uniform lattice in a higher rank non-compact connected semi-simple Lie group. In Subsection \ref{raag}, we prove that any non-join right-angled Artin group has an algebraic rank of $1$. In Section \ref{relhypgppre}, we study algebraic rank of groups which act on $\mathrm{CAT}(0)$ spaces with isolated flats. More precisely, if a group $G$ acts geometrically on $\mathrm{CAT}(0)$ space with isolated flats $\mathcal{F}$ and $|\mathcal{F}| \neq 1$, then $\mathrm{rank}(G) \leq 1$. It is well known that such a group $G$ is hyperbolic relative to a family of stabilizers of flats in $\mathcal{F}$. We use the dynamics of a relatively hyperbolic group acting on the boundary of a $\delta$-hyperbolic space to prove that the algebraic rank of a relative hyperbolic group is $\leq 1$ if there are at least two peripheral subgroups containing elements of infinite order. It follows immediately that $\mathrm{rank}(G) \leq 1$. This paper is part of author's Ph.D. thesis. The author thanks the thesis advisor Jean Lafont for his guidance throughout this research project. The author also thanks Mike Davis for helpful conversations concerning Coxeter groups. \section{Algebraic Rank of Groups}\label{algrk} \begin{definition} For a given group $G$, let $\mathcal{A}_{i}(G)$ be the set consisting of elements such that the centralizer contains a free abelian subgroup of rank $\leq i$ as a subgroup of finite index. Define $r(G)$ to be the minimum $i$ such that $G$ can be expressed as the union of finitely many translates of $\mathcal{A}_{i}(G)$. In other words, $$\displaystyle{r(G) = min\{i \,|\,\, \textrm{there are finitely many elements}\, g_{j} \in G \, \textrm{such that}\,\, G = \bigcup_{j} g_{j}\mathcal{A}_{i}(G)\}}$$ Finally, the \emph{algebraic rank} of $G$, $\mathrm{rank}(G)$, is the supremum of $r(G^{*})$ over all finite index subgroups $G^{*}$ of $G$. \end{definition} We allow the possibility that $\mathrm{rank}(G)=0$. For example, if $G$ is a finite group, then $\mathrm{rank}(G)=0$. On the other hand, if $G$ is torsion-free, then $\mathrm{rank}(G) >0$ : suppose that $\mathrm{rank}(G)=0$. In particular, $r(G)=0$. Then the set $\mathcal{A}_{0}(G)$ must be non-empty. But $\mathcal{A}_{0}(G)$ is a subset of the set of finite order elements in $G$. We set $\mathrm{rank}(G)=\infty$ if the sets $\mathcal{A}_{i}(G)$ are empty, or if $G$ cannot be covered by finitely many translates of any of the sets $\mathcal{A}_{i}(G)$. For example, if a group $G$ has an infinitely generated free abelian center, then $\mathrm{rank}(G)=\infty$. In fact, there exist finitely presented examples of such groups. More specifically, Hall obtained in {\cite{Hal}} the existence of a finitely generated group which has infinitely generated free abelian center. Using {\cite{Oul}}, we can obtain a finitely presented group having infinitely generated free abelian center. \begin{remark} $r(G)$ is not necessarily equal to $\mathrm{rank}(G)$. Following \cite{BalEbe}, we present an example of a group satisfying $r(G) < \mathrm{rank}(G)$. See {\cite[Section 4]{BalEbe}} for more examples. Let $G$ be the fundamental group of a flat Klein bottle, acting on $\mathbb{E}^{2}$ by isometries. (In the simplest case $G$ is generated by $\phi_{1} : (x,y) \to (x+1,-y)$ and $\phi_{2}:(x,y) \to (x,y+1)$.) The set $\mathcal{A}_{1}(G)$ consists of all elements of $G$ that reverse the orientation of $\mathbb{E}^{2}$. Then $G=\mathcal{A}_{1}(G) \cup \gamma\mathcal{A}_{1}(G)$, for any $\gamma \in \mathcal{A}_{1}(G)$. Therefore, $r(G) < 2=\mathrm{rank}(G) = \mathrm{rank}(\mathbb{E}^{2})$. \end{remark} We close the section by mentioning that algebraic rank of groups behaves well under products and taking finite index subgroups. \begin{proposition}{\cite[Proposition 2.1]{BalEbe}}\label{algrkpro} Let $G$ be an abstract group. \begin{enumerate} \item If $G'$ is a finite index subgroup of $G$, then $r(G) \leq r(G')$ and $\mathrm{rank}(G)=\mathrm{rank}(G')$. \item If $G=G_1 \times \cdots \times G_n$, then $$r(G) = \sum_{i=1}^{n} r(G_{i}),\qquad \mathrm{rank}(G) = \sum_{i=1}^{n} \mathrm{rank}(G_{i})$$ \end{enumerate} \end{proposition} \section{Right -Angled Coxeter Groups and Artin Groups} \subsection{Coxeter Groups} A \emph{Coxeter system} $(W,S)$ is a group $W$ and a set $S=\{s_1, s_2, \cdots\}$ of generators such that $W$ has the following presentation $$W=\langle S\, |\, (s_{i}s_{j})^{m_{ij}} =1, s_{i},s_{j} \in S\rangle,$$ where $m_{ii}=1$ and if $i \neq j$, then $m_{ij}=m_{ji}$ is a positive integer $\geq 2$ or $\infty$ (in which case we omit the relation between $s_{i}$ and $s_{j}$). $W$ is called a \emph{Coxeter group}. A Coxeter system $(W,S)$ is called \emph{irreducible} if $S$ cannot be partitioned into two nonempty disjoint subsets $S'$ and $S''$ such that each element in $S'$ commutes with each element in $S''$. The cardinality $|S|$ of $S$ is called the \emph{rank} of $W$ and we assume that $|S|$ is finite in this section. A Coxeter group $W$ is \emph{spherical} if $W$ is finite and \emph{affine} if $W$ has a finite index free abelian subgroup. For any subset $J \subset S$, we denote by $W_{J}$ the subgroup of $W$ generated by $J$. We call $W_{J}$ \emph{a standard parabolic subgroup}, and any conjugate of a standard parabolic subgroup is called \emph{a parabolic subgroup}. For any subset $A \subset W$, the \emph{parabolic closure} $\mathrm{Pc}(A)$ of $A$ is the smallest parabolic subgroup containing $A$. An element $\gamma$ is called \emph{essential} if $\mathrm{Pc}(\gamma)=W$. Associated to any Coxeter group $W$, there is a $\mathrm{CAT(0)}$ polyhedral cell complex $\Sigma_{W}$, which is called the \emph{Davis complex}, upon which $W$ acts properly discontinuously and cocompactly by isometries. $\Sigma_{W}$ can be cellulated by so called \emph{Coxeter polytopes} and, with a natural Euclidean metric on each Coxeter polytope, inherits a piecewise Euclidean metric. It was Gromov\,(right-angled case, \cite{Gro}) and Moussong\,(general case, \cite{Mou}), who showed that $\Sigma_{W}$, with this metric, is $\mathrm{CAT}(0)$. See \cite{DavBOOK} for details. An element $\gamma \in W$ is said to have \emph{rank one} if it is hyperbolic and if some (and hence any) of its axes in $\Sigma_{W}$ has rank one. In \cite{CapFuj}, Caprace and Fujiwara study rank one elements in Coxeter groups. In particular, an element $\gamma$ has rank one if and only if its centralizer is virtually infinite cyclic. In other words, any rank one element is contained in $\mathcal{A}_{1}(W)$. Right-angled Coxeter groups are Coxeter groups for which $m_{ij}=2$ or $\infty$ for $i \neq j$. In this case, the Davis complex $\Sigma_{W}$ is a $\mathrm{CAT}(0)$ cubical complex. In this subsection, we prove that any infinite irreducible non-affine right-angled Coxeter group has an algebraic rank of $1$. \begin{remark} \begin{enumerate} \item Any spherical Coxeter group has an algebraic rank of $0$. (See Section \ref{algrk}.) \item It is a consequence of Selberg's lemma that every infinite Coxeter group $W$ has a torsion-free subgroup of finite index. Therefore, such groups satisfy ${\mathrm{rank}}(W) \geq 1$. \item If $W$ is infinite, irreducible and affine, then ${\mathrm{rank}}(W) = |S|-1$. \item Suppose that $W$ is infinite and reducible, $W=W_{T_{1}} \times W_{T_{2}} \times \cdots \times W_{T_{n}}$. By Proposition \ref{algrkpro}, $\displaystyle{{\mathrm{rank}}(W) =\sum_{i=1}^{n} {\mathrm{rank}}(W_{T_{i}})}$. Therefore, ${\mathrm{rank}}(W) > 1$ unless $W$ is virtually cyclic. \end{enumerate} \end{remark} Hereafter, we assume that $W$ is an infinite irreducible non-affine right-angled Coxeter group. Tits' solution to the word problem for Coxeter groups states that any two reduced expressions represent the same element in $W$ if and only if one can be transformed into the other by a series of replacements of the alternating subword $st$ by the subword $ts$. (See \cite[Sec. 3.4]{DavBOOK}.) This implies \begin{lemma}\label{word} For $w \in W$, let $S(w)$ be the set of generators appearing in some (and hence any) reduced expression for $w$. If $s \in S(w)$ appears an odd (respectively, even) number of times in some expression for $w$, then $s$ appears an odd (respectively, even) number of times in any expression for $w$. \end{lemma} Let $\mathcal{H}$ be the set of rank one elements in $W$. It is not difficult to find rank one elements in $W$. For example, any essential element in $W$ has virtually infinite cyclic centralizer, therefore it has rank one. (See \cite[Corollary 6.3.10]{Kra}) \begin{lemma} Let $w$ be an element such that some (and hence any) reduced expression for $w$ has the following property : all generators appear, and each generator appears an odd number of times. Then $w$ is essential, and hence, $w$ has rank one. \end{lemma} \begin{proof} Suppose that $\mathrm{Pc}(w) = uW_{J}u^{-1}$ for some $u$ and some $J \subset S$. Suppose that $s \notin J$. Then any reduced expression for words in $uW_{J}u^{-1}$ contains $s$ an even number of times. Lemma \ref{word} gives contradiction and we can conclude that $s \in J$. This proves that $J=S$. \end{proof} \begin{proposition}\label{coxrk} $r(W) \leq 1$. \end{proposition} \begin{proof} Recall $\mathcal{H} \subset \mathcal{A}_{1}(W)$. Let $\mathcal{S} =\{s_{1}\cdots s_{k} \,| \, s_{i} \in S, \mathrm{distinct}, k \leq n\}$, where $n=|S|$. In other words, $\mathcal{S}$ is the set of all possible products of distinct generators. We prove that for any element $t \in W \setminus \mathcal{H}$, there exists $g \in \mathcal{S}$ such that $gt \in \mathcal{H}$. Let $t \in W \setminus \mathcal{H}$ be given and consider any reduced expression $\mathbf{t}$ for $t$. Multiply $\mathbf{t}$ by all generators appearing an even (including zero) number of times in $\mathbf{t}$. $$(s_{i_{1}}\cdots s_{i_{n}})\mathbf{t}$$ Then the resulting word, and hence any reduced expression, has the property that all generators appear and each generator appears an odd number of times. Therefore, the element represented by this word is essential, and hence, it has rank one. $$\displaystyle{W=\mathcal{H} \bigcup (\bigcup_{\alpha \in \mathcal{S}} \alpha \mathcal{H}}).$$ This proves that $r(W) \leq 1$. \end{proof} In order to prove that ${\mathrm{rank}}(W) =1$, we need to show $r(T) \leq 1$ for any finite index subgroup $T$ of $W$. But it seems that the above argument does not work for $T$. Because, in the proof of Proposition \ref{coxrk}, $(s_{i_{1}}\cdots s_{i_{n}})$ does not necessarily represent an element in $T$. Once one takes powers on $(s_{i_{1}}\cdots s_{i_{n}})$ to get an element in $T$, the argument fails to apply. In particular, if the index $[W:T]$ is even, all generators in $(s_{i_{1}}\cdots s_{i_{n}})^{[W:T]}$ appear an even number of times. As an example, one can consider the commutator subgroup of $W$. Since the commutator subgroup misses all all-odd elements, it does not contain elements of type appeared in Lemma 7. Therefore, we take a different approach to prove $r(T) \leq 1$. \begin{definition} Let $\mathbf{w}$ be a reduced word in $S$. For any generator $s$ appearing in $\mathbf{w}$, let $$\mathbf{w}=\mathbf{w_{0}}s \mathbf{w_{1}} s \cdots s \mathbf{w_{k}} s \mathbf{w_{k+1}},$$ where $\mathbf{w_{i}}$ does not contain $s$ for all $0 \leq i \leq k+1$. Note that $\mathbf{w_0}$ and $\mathbf{w_{k+1}}$ are allowed to be empty, but each $\mathbf{w_{i}} \neq \emptyset$ for $1 \leq i \leq k$. \begin{enumerate} \item $\mathbf{w}$ is said to be \emph{$s$-minimal} if each subword $\mathbf{w_{i}}$, for $1 \leq i \leq k$, contains a $s$-blocker, i.e., a generator $s' \in S$ such that $ss' \neq s's$. We consider $\mathbf{w}$ to be vacuously $s$-minimal if $s$ appears only once in $\mathbf{w}$. \item $\mathbf{w}$ is said to be \emph{$s$-good} if $\mathbf{w}$ is $s$-minimal and $\mathbf{w_{k+1}w_{0}}$ contains a $s$-blocker for $k \geq 1$. In the case that $k=0$, $\mathbf{w}$ is considered to be $s$-good. \end{enumerate} \end{definition} \begin{remark} Any reduced word $\mathbf{w}$ is $s$-minimal for all generators $s$ appearing in $\mathbf{w}$. For a generator $s$ appearing in $\mathbf{w}$, let $$\mathbf{w}=\mathbf{w_{0}}s \mathbf{w_{1}} s \cdots s \mathbf{w_{k}} s \mathbf{w_{k+1}}.$$ If $\mathbf{w_{i}}$ does not contain a $s$-blocker for some $1 \leq i \leq k$, then $s \mathbf{w_{i}} s = \mathbf{w_{i}}$, which is a contradiction. \end{remark} \begin{lemma} Let $\mathbf{w}$ be a reduced word which is $s$-good for all $s \in S$. Then the element represented by $\mathbf{w}$ is essential. \end{lemma} \begin{proof} Suppose that $s$ appears once in $\mathbf{w}$. In other words, $\mathbf{w} = \mathbf{w_{0}}s\mathbf{w_{1}}$. Suppose the element represented by $\mathbf{w}$ is in $u^{-1}W_{J}u$ for some $u \in W$ and $J \subset S$. Then the element represented by $\mathbf{u}\mathbf{w}\mathbf{u}^{-1}$ lies in $W_{J}$, where $\mathbf{u}$ is any reduced expression for $u$. In some (any) reduced expression of $\mathbf{u}\mathbf{w}\mathbf{u}^{-1}$, $s$ appears an odd number of times. In particular, $s$ appears. Therefore, $s \in J$. Suppose $s$ appears at least twice in $\mathbf{w}$. In other words, $$\mathbf{w}=\mathbf{w_{0}}s \mathbf{w_{1}} s \cdots s \mathbf{w_{k}} s \mathbf{w_{k+1}},$$ for $k \geq 1$. Suppose that the element represented by $\mathbf{w}$ is in $u^{-1}W_{J}u$ for some $u \in W$ and $J \subset S$. Then the element represented by $$\mathbf{u}\mathbf{w}\mathbf{u}^{-1} = \mathbf{u}\mathbf{w_{0}}s \mathbf{w_{1}} s \cdots s \mathbf{w_{k}} s \mathbf{w_{k+1}}\mathbf{u}^{-1}$$ lies in $W_{J}$, where $\mathbf{u}$ is any reduced expression of $u$. Assume that $s \notin J$. Then $\mathbf{u}$ must contain $s$. We prove that one of two $s$'s in $\mathbf{u}$ and $\mathbf{u}^{-1}$ cannot be cancelled off. By way of contradiction, let us assume that both can be cancelled. Since $\mathbf{w}$ is $s$-good, there exists at least one $s$-blocker in $\mathbf{w_{0}}$ or $\mathbf{w_{k+1}}$. Without loss of generality, we assume that $s$-blocker lies in $\mathbf{w_{0}}$. (A symmetric argument applies if it lies in $\mathbf{w_{k+1}}$.) Take the first $s$-blocker in $\mathbf{w_0}$ and call it $s_1$. It follows that $\mathbf{u}$ must contain $s_1$ and the last $s_{1}$ occurs after the last $s$ in $\mathbf{u}$. $$\mathbf{u}\mathbf{w}\mathbf{u}^{-1} = (\cdots s \cdots s_1 \cdots)(\cdots s_1 \cdots) s \mathbf{w_{1}} s \cdots s \mathbf{w_{k}} s \mathbf{w_{k+1}}(\cdots s_{1} \cdots s \cdots)$$ In order for the last $s$ to be cancelled off, $s_1$ must occur in $\mathbf{w_{k+1}}$. $$\mathbf{u}\mathbf{w}\mathbf{u}^{-1} = (\cdots s \cdots s_1 \cdots)(\cdots s_1 \cdots) s \mathbf{w_{1}} s \cdots s \mathbf{w_{k}} s (\cdots s_1 \cdots)(\cdots s_{1} \cdots s \cdots),$$ where the $s_1 \in \mathbf{w_{k+1}}$ written above is the last occurrence of $s_1$ in $\mathbf{w_{k+1}}$. So before being able to cancel the first and the last $s$, we need to cancel out the intermediate blocker $s_1$. Now $\mathbf{w}$ is $s_1$-good. Therefore, there exists an $s_1$-blocker on the left of the first $s_1 \in \mathbf{w_{0}}$ or on the right of the last $s_1 \in \mathbf{w_{k+1}}$. Take the first $s_1$-blocker in $\mathbf{w_{0}}$ or the last $s_1$-blocker in $\mathbf{w_{k+1}}$ and call it $s_2$. Note that $s_2 \neq s$ and before canceling the $s_1$, we must first be able to cancel out the $s_1$-blocker $s_2$. As in the last paragraph, this forces $\mathbf{u}\mathbf{w}\mathbf{u}^{-1}$ to be of the form $$(\cdots s_1 \cdots s_2 \cdots)(\cdots s_2 \cdots s_1 \cdots) s \mathbf{w_{1}} s \cdots s \mathbf{w_{k}} s (\cdots s_1 \cdots s_2 \cdots)(\cdots s_2 \cdots s_{1} \cdots)$$ Note that $\mathbf{w}$ is $s_2$-good, and hence, there exists an $s_2$-blocker $(\neq s, s_1)$ on the left of the first $s_2$ or on the right of the last $s_2$ in $\mathbf{w}$. But since $|S| < \infty$, this process must stop in finitely many stages, which proves that one of $s$'s in $\mathbf{u}$ and $\mathbf{u}^{-1}$ cannot be cancelled off. Therefore, $s \in J$. The element represented by $\mathbf{w}$ is essential. \end{proof} Let $T$ be a proper finite index subgroup of $W$. Assume that $T$ is normal and let $n=[W:T] \geq 2$. In order to prove that $r(T) \leq 1$, we need to consider two types of generators for a given reduced word $\mathbf{w}$ representing an element in $T$ : (1) a generator $s$ does not appear in $\mathbf{w}$ and (2) a generator $s$ appears, but $\mathbf{w}$ is not $s$-good. We begin with generators of type (1). Let $\mathbf{w}$ be a reduced word in $S$ representing an element in $T$ and assume that $\mathbf{w}$ misses a generator $s$. Choose $s' \in S$ such that $ss' \neq s's$. Note that such $s'$ always exists because $W$ is infinite irreducible. Next choose $s'' \in S$ such that either $ss'' \neq s''s$ or $s's'' \neq s''s'$. Note that such $s''$ always exists, otherwise $W=W_{\{s,s'\}} \times W_{S\setminus\{s,s'\}}$, contradicting irreducibility (if $|S| >2$) or non-affine (if $|S| = 2$). \begin{lemma}\label{coxlemma1} \begin{enumerate} \item Suppose that $ss''\neq s''s$. Then any reduced expression $\mathbf{r}$ of $(s''ss')^{n}\mathbf{w}$ has the following property: \begin{enumerate} \item The element represented by $\mathbf{r}$ is in $T$. This is obvious. \item $\mathbf{r}$ is $s$-good. \item $\mathbf{r}$ is $s'$-good. \item For $t \neq s,s',s''$, if $\mathbf{w}$ is $t$-good, then $\mathbf{r}$ is also $t$-good. \end{enumerate} \item Suppose that $s's'' \neq s''s'$. Then any reduced expression $\mathbf{r'}$ of $(s's''ss's'')^{n}\mathbf{w}$ has the following property: \begin{enumerate} \item The element represented by $\mathbf{r'}$ is in $T$. \item $\mathbf{r'}$ is $s$-good. \item $\mathbf{r'}$ is $s''$-good. \item For $t \neq s,s',s''$, if $\mathbf{w}$ is $t$-good, then $\mathbf{r'}$ is also $t$-good. \end{enumerate} \end{enumerate} \end{lemma} \begin{proof} We prove the first statement only. The second statement can be proved by exactly the same argument as the first. Consider $$(s''ss')^{n}\mathbf{w}=(s''ss')(s''ss')\cdots(s''ss')\mathbf{w}=(s'')s(s's'')s(s'\cdots s'')s(s'\mathbf{w}).$$ Since $ss' \neq s's$, $ss''\neq s''s$, $s'$ or $s''$ before the last occurrence of $s$ cannot be cancelled. Since $s$ does not appear in $\mathbf{w}$, $\mathbf{r}$ is $s$-good. Let $\mathbf{w}=\mathbf{w'_{0}}s'\mathbf{w'_{1}}s'\cdots \mathbf{w'_{l}}s' \mathbf{w'_{l+1}}$ and consider $$(s''ss')^{n}\mathbf{w} = (s''s)s'(s''s)s'\cdots s'(s''s)s'\mathbf{w'_{0}}s'\mathbf{w'_{1}}s'\cdots \mathbf{w'_{l}}s' \mathbf{w'_{l+1}}.$$ Note that if the subword $\mathbf{w'_{0}}$ does not contain $s'$-blocker, then $s'\mathbf{w'_{0}}s'$ is reduced to $\mathbf{w'_{0}}$. But $s$ is an $s'$-blocker. Therefore, there exists at least one $s'$-blocker in the reduced expression of $(s''ss')\mathbf{w'_{0}}s'\mathbf{w'_{1}}$, even if $\mathbf{w'_{0}}$ does not contain $s'$-blocker. (Recall that the letter $s$ does not appear in $\mathbf{w'_{0}}$ and $\mathbf{w'_{1}}$, so it cannot be cancelled off.) Suppose that $\mathbf{w}$ is $t$-good for $t \neq s,s',s''$. If neither $s,s'$ or $s''$ is $t$-blocker, then $\mathbf{r}$ is obviously $t$-good. Consider the case that either $s,s'$ or $s''$ is a $t$-blocker. $$(s''ss')^{n}\mathbf{w}=(s''ss')(s''ss')\cdots(s''ss')\mathbf{w''_{0}}t\mathbf{w''_{1}}t\cdots \mathbf{w''_{m}}t \mathbf{w''_{m+1}}$$ Note that $s$ is not in $\mathbf{w}$ and $n \geq 2$. Therefore, there exists at least one $t$-blocker in the reduced expression of $(s''ss')(s''ss')\cdots(s''ss')\mathbf{w''_{0}}$. It follows that $\mathbf{r}$ is $t$-good. \end{proof} \begin{remark}\label{allgen} \begin{enumerate} \item In Lemma \ref{coxlemma1}, $\mathbf{r}$ is not necessarily $s''$-good. Similarly, $\mathbf{r'}$ is not necessarily $s'$-good. Therefore, the number of good generators of $\mathbf{r}$ or $\mathbf{r'}$ might be equal to the number of good generators of $\mathbf{w}$. But note that $s$ appears in $\mathbf{r}$ and $\mathbf{r'}$, and all letters appearing in $\mathbf{w}$ still appear in $\mathbf{r}$ and $\mathbf{r'}$. \item Let $\mathbf{w}$ be a reduced word in $S$ representing an element in $T$. By multiplying words as in Lemma \ref{coxlemma1}, we can obtain $\mathbf{w':=y_{k}y_{k-1}\cdots y_{1}w}$ such that the element represented by $\mathbf{w'}$ is in $T$ and all generators appear in $\mathbf{w'}$. \item There exists a finite set $\mathcal{R}$ of words such that for any reduced word $\mathbf{w}$ representing an element in $T$, there exists some $\mathbf{r} \in \mathcal{R}$ for which $\mathbf{rw}$ represents an element in $T$ and all generators appear in $\mathbf{rw}$ . \end{enumerate} \end{remark} Next, we consider generators of type (2). \begin{definition} Let $\mathbf{w}$ be a reduced word in $S$ such that all generators appear. Define $B(\mathbf{w})$ be the set of generators for which $\mathbf{w}$ is not good, i.e., $B(\mathbf{w})$ consists of all the ``bad" generators. For $B \subset S$, a word $\mathbf{v}$ is called a \emph{$B$-cancellator} if, for any $\mathbf{w}$ such that $B=B(\mathbf{w})$, any reduced expression of $\mathbf{v}\mathbf{w}$ is $s$-good for all $s \in S$. \end{definition} The following lemma tells us that $B$-cancellators exist and can be chosen to represent an element in $T$. Note that there are only finitely many subsets of $S$. Therefore, we can form finitely many cancellators. Let $\mathbf{w}$ be a reduced word in $S$ such that all generators appear and $\mathbf{w}$ represents an element in $T$. Suppose that $\mathbf{w}$ is not $s$-good. Choose $s' \in S$ such that $ss' \neq s's$. Note that such $s'$ always exists because $W$ is infinite irreducible. Also we choose $s'' \in S$ such that either $ss'' \neq s''s$ or $s's'' \neq s''s'$. Note that such $s''$ always exists, otherwise $W=W_{\{s,s'\}} \times W_{S\setminus\{s,s'\}}$. \begin{lemma}\label{coxlemma2} \begin{enumerate} \item Suppose that $ss''\neq s''s$. Then any reduced expression $\mathbf{r}$ of $(s''ss')^{n}\mathbf{w}$ has the following property: \begin{enumerate} \item The element represented by $\mathbf{r}$ is in $T$. This is obvious. \item $\mathbf{r}$ is $s$-good. \item $\mathbf{r}$ is $s'$-good. \item $\mathbf{r}$ is $s''$-good. \item For $t \neq s,s',s''$, if $\mathbf{w}$ is $t$-good, then $\mathbf{r}$ is also $t$-good. \end{enumerate} \item Suppose that $s's'' \neq s''s'$. Then any reduced expression $\mathbf{r'}$ of $(s's''ss's'')^{n}\mathbf{w}$ has the following property: \begin{enumerate} \item The element represented by $\mathbf{r'}$ is in $T$. \item $\mathbf{r'}$ is $s$-good. \item $\mathbf{r'}$ is $s'$-good. \item $\mathbf{r'}$ is $s''$-good. \item For $t \neq s,s',s''$, if $\mathbf{w}$ is $t$-good, then $\mathbf{r'}$ is also $t$-good. \end{enumerate} \end{enumerate} \end{lemma} \begin{proof} Again, we prove the first statement only. The second statement can be proved by exactly the same argument as the first. Let $\mathbf{w} = \mathbf{w_{0}}s \mathbf{w_{1}} s \cdots s \mathbf{w_{k}} s \mathbf{w_{k+1}}$ and consider $$(s''ss')^{n}\mathbf{w} = (s'')s(s's'')\cdots(s's'')s(s'\mathbf{w_{0}})s \mathbf{w_{1}} s \cdots s \mathbf{w_{k}} s \mathbf{w_{k+1}}.$$ In the reduced expression of $ss'\mathbf{w_{0}}s$, since $s$ and $s'$ don't commute, $s'$ is $s$-blocker. (Note that $\mathbf{w_{0}}$ does not contain $s'$, since, by assumption, $s \in B(\mathbf{w}))$. Also the first $s''$ is an $s$-blocker. Hence $\mathbf{r}$ is $s$-good. $$(s''ss')^{n}\mathbf{w}=(s''s)s'(s''s)s'\cdots(s''s)s'\mathbf{w'_{0}}s'\mathbf{w'_{1}}s'\cdots \mathbf{w'_{m}}s' \mathbf{w'_{j+1}}$$ Secondly, note that $s$ and $s'$ don't commute. Since $\mathbf{w}$ is not $s$-good, the first $s'$ in $\mathbf{w}$ should occur after the first $s$ in $\mathbf{w}$, i.e., $s \in \mathbf{w'_{0}}$. It follows that $s$ is an $s'$-blocker in $\mathbf{w'_{0}}$. Hence $\mathbf{r}$ is $s'$-good. $$(s''ss')^{n}\mathbf{w}=s''(ss')s''(ss')\cdots s''(ss'\mathbf{w''_{0}})s''\mathbf{w''_{1}}s''\cdots \mathbf{w''_{i}}s'' \mathbf{w''_{i+1}}$$ Note that $s$ and $s''$ don't commute. Since $\mathbf{w}$ is not $s$-good, the last $s''$ in $\mathbf{w}$ should occur before the last $s$ in $\mathbf{w}$, i.e., $s \in \mathbf{w''_{i+1}}$. It follows that $\mathbf{w''_{i+1}}$ contains an $s''$-blocker $s$. $\mathbf{r}$ is $s''$-good. Suppose that $\mathbf{w}$ is $t$-good for $t \neq s, s', s''$. If neither $s,s'$ or $s''$ is $t$-blocker, then $\mathbf{r}$ is $t$-good. Let $$(s''ss')^{n}\mathbf{w}=(s''ss')(s''ss')\cdots(s''ss')\mathbf{w'''_{0}}t\mathbf{w'''_{1}}t\cdots \mathbf{w'''_{h}}t \mathbf{w'''_{h+1}}.$$ If $s$ is a $t$-blocker, the first $t$ occur after the first $s$ in $\mathbf{w}$, i.e., $s \in \mathbf{w'''_{0}}$. Furthermore, this $s$ cannot be cancelled off, because $s$ and $s'$ don't commute. It follows that there exists at least one $t$-blocker in the reduced expression of $(s''ss')\mathbf{w'''_{0}}$. $\mathbf{r}$ is $t$-good. Next, suppose that $s'$ or $s''$ is $t$-blocker. The $s$ in the last $(s''ss')$ cannot be cancelled off. Therefore, there exists at least one $t$-blocker in the reduced expression of $(s''ss')(s''ss')\cdots(s''ss')\mathbf{w'''_{0}}$. $\mathbf{r}$ is $t$-good. \end{proof} \begin{corollary}\label{cancelator} For a given $B \subset S$, a $B$-cancellator $\mathbf{v_{B}}$ exists and can be chosen to represent an element in $T$. \end{corollary} \begin{proof} Let $\mathbf{w}$ be a reduced word in $S$, with all generators appearing, such that $B(\mathbf{w})=B$. Choose a generator $s_{1} \in B(\mathbf{w})$. Apply the lemma to obtain a word $\mathbf{v_{1}}$ representing an element in $T$ such that any reduced expression $\mathbf{r}_{1}$ of $\mathbf{v_{1}w}$ is $s_{1}$-good. Consider $\mathbf{r}_{1}$. From Lemma \ref{coxlemma2}, $B(\mathbf{r}_{1}) \subset B(\mathbf{w})$ and $|B(\mathbf{r}_{1})| < |B(\mathbf{w})|$. Choose a generator $s_{2} \in B(\mathbf{r}_{1})$ and apply the lemma to obtain a word $\mathbf{v_{2}}$ representing an element in $T$ such that any reduced expression $\mathbf{r}_{2}$ of $\mathbf{v_{2}}\mathbf{r}_{1}$ is $s_{2}$-good. Continuing this process, at most $|B(\mathbf{w})|$ number of times, we obtain a word $\mathbf{v_{k}}\mathbf{v_{k-1}}\cdots \mathbf{v_{1}} \mathbf{w}, k \leq |B(\mathbf{w})|$ whose reduced expression is $s$-good for all $s \in S$. Such a $\mathbf{v_{k}}\mathbf{v_{k-1}}\cdots \mathbf{v_{1}}$ is the desired $B$-cancellator. \end{proof} \begin{corollary}\label{norcox} $r(T) \leq 1$. \end{corollary} \begin{proof} Let $\mathcal{H}_{T} = \mathcal{H} \bigcap T$. For $g \in \mathcal{H}_{T}$, the centralizer $C_{W}(g)$ in $W$ is virtually infinite cyclic, and hence, $C_{T}(g)$ is also virtually infinite cyclic. It follows that $\mathcal{H}_{T} \subset \mathcal{A}_{1}(T)$. Let $g' \in T \setminus \mathcal{H}_{T}$ be given. By Remark \ref{allgen} and Corollary \ref{cancelator}, we can find a $\mathbf{r} \in \mathcal{R}$ and a cancellator $\mathbf{v}$ such that any reduced expression of $\mathbf{v}\mathbf{rw'}$ is $s$-good for all $s \in S$, where $\mathbf{w'}$ is any reduced expression for $g'$. Note that $\mathbf{v}\mathbf{rw'}$ represents an element in $\mathcal{H}_{T}$. Therefore, $$T = \bigcup_{B \subseteq S, \mathbf{r} \in \mathcal{R}} \mathbf{r}^{-1}\mathbf{v_B}^{-1}\mathcal{H}_{T},$$ where each $\mathbf{v}_{B}$ is a $B$-cancellator. \end{proof} \begin{proposition}\label{cox} Let $W$ be an infinite, irreducible, and non-affine right-angled Coxeter group. Then ${\mathrm{rank}}(W)=1$. \end{proposition} \begin{proof} By Proposition \ref{coxrk}, it suffices to show that $r(W') \leq 1$ for any finite index subgroup $W'$ of $W$. By taking the normal core, we obtain a finite index normal subgroup $W''$ of $W$ and, by Corollary \ref{norcox}, $r(W'') \leq 1$. Finally, Proposition \ref{algrkpro} implies that $r(W') \leq 1$. \end{proof} We close the subsection by introducing two corollaries of Proposition \ref{cox}. Two groups $G_1$ and $G_2$ are said to be \emph{commensurable} if there exist $H_{i}$, for $i=1,2$, such that $[G_{i}:H_{i}] < \infty $ or $[H_{i}:G_{i}] < \infty$ and $H_{1}$ is isomorphic to $H_{2}$. Davis proved that any infinite irreducible non-affine Coxeter group $W$ cannot be a uniform lattice $\Gamma$ in a higher rank non-compact connected semi-simple Lie group (See \cite[Corollary 10.9.8]{DavBOOK}) and conjectured $W$ cannot be commensurable to $\Gamma$. (See \cite{Dav}) The conjecture is known to be true, see for example, \cite{CooLon}, \cite{Gon}, \cite{Sin}. Proposition \ref{cox} provides a new proof of the conjecture for right-angled Coxeter groups. \begin{corollary}\label{commen} Let $W$ be an infinite irreducible non-affine right-angled Coxeter group. Then $W$ is not commensurable to any uniform lattice in a higher rank non-compact connected semi-simple Lie group $G$. \end{corollary} \begin{proof} Let $\Lambda$ be a uniform lattice in $G$. By \cite[Theorem 3.11]{BalEbe}, ${\mathrm{rank}}(\Lambda) = {\mathrm{rank}}(G) \geq 2$. Applying Proposition \ref{algrkpro}, we obtain that any group $\Gamma$ commensurable to $\Lambda$ satisfies ${\mathrm{rank}}(\Gamma) = {\mathrm{rank}}(\Lambda) \geq 2$. On the other hand, for any finite index subgroup $W'$ of $W$, ${\mathrm{rank}}(W)={\mathrm{rank}}(W')=1$. Therefore, $W$ and $\Lambda$ cannot be commensurable. \end{proof} The other corollary follows from Quasi-Isometry rigidity theorem due to Kleiner-Leeb (\cite{KleLee}) and Eskin-Farb (\cite{EskFar}). \begin{corollary} Let $W$ be an infinite irreducible non-affine right-angled Coxeter group. Then $W$ is not quasi-isometric to any uniform lattice in a higher rank non-compact connected semi-simple Lie group $G$. \end{corollary} \begin{proof} By QI-rigidity theorem, if $W$ is quasi-isometric to a uniform lattice in a higher rank non-compact connected semi-simple Lie group, $W$ should be commensurable to a lattice. By Corollary \ref{commen}, $W$ cannot be commensurable to the lattice. \end{proof} \subsection{Algebraic Rank of Right-Angled Artin Groups}\label{raag} An Artin group $A$ is a group with the following presentation : $$A=\langle s_{1}, \cdots, s_{n} | \underbrace{s_{i}s_{j}\cdots}_{m_{ij}} = \underbrace{s_{j}s_{i}\cdots}_{m_{ij}}, \mathrm{\,\,for\,\,all\,\,} i \neq j \rangle,$$ where $m_{ij}=m_{ji}$ is an integer $\geq 2$ or $m_{ij}=\infty$ in which case we omit the relation between $s_{i}$ and $s_{j}$. As one can see, by adding relations $s_{i}=s_{i}^{-1}$ to the presentation, we obtain a Coxeter group. Right-angled Artin groups are those Artin groups for which all $m_{ij}=2$ or $\infty$ for $i \neq j$. One of the easy ways of defining a right-angled Coxeter group or a right-angled Artin group is via \emph{the defining graph} $\Gamma$. This is the graph whose vertices are labeled by $S=\{s_{1}, \cdots, s_{n}\}$ and two vertices $s_{i}$ and $s_{j}$ are connected if $m_{ij}=2$. We denote by $A_{\Gamma}\, (W_{\Gamma}$, respectively) the right-angled Artin group (the right-angled Coxeter group, respectively) associated to a finite simplicial graph $\Gamma$. For example, if $\Gamma$ consists of $n$ vertices and no edges, then $A_{\Gamma}$ is the free group on $n$ generators. At the other extreme, if $\Gamma$ is a complete graph with $n$ vertices, $A_{\Gamma}$ is the free abelian group of rank $n$. Analogous to the Coxeter group situation, there is a $\mathrm{CAT}(0)$ space associated to a right-angled Artin group $A_{\Gamma}$, which can be constructed by the following process : begin with a wedge of circles attached to a point $x_{0}$ and labeled by the generators $s_{1}, \cdots s_{n}$. For each edge connecting $s_{i}$ and $s_{j}$ in $\Gamma$, attach a $2$-torus with boundary labeled by the relator $s_{i}s_{j}s_{i}^{-1}s_{j}^{-1}$. For each triangle connecting $s_{i},s_{j},s_{k}$ in $\Gamma$, attach a $3$-torus with faces corresponding to the tori for the three edges of triangle. Continuing this process, attach a $k$-torus for each set of $k$-mutually commuting generators. The resulting cube complex is called \emph{a Salvetti complex} for $A_{\Gamma}$ and denoted by $\mathcal{S}_{\Gamma}$. It is easy to verify that the fundamental group of $\mathcal{S}_{\Gamma}$ is $A_{\Gamma}$ and the link of the unique vertex $x_{0}$ is a flag. It follows from Gromov's criterion that the universal cover $X_{\Gamma}$ of the complex $\mathcal{S}_{\Gamma}$ is a $\mathrm{CAT}(0)$ cube complex, and $A_{\Gamma}$ acts on $X_{\Gamma}$ freely and cocompactly. Given two graphs $\Gamma_{1}, \Gamma_{2}$, their \emph{join} is the graph obtained by connecting every vertex of $\Gamma_{1}$ to every vertex of $\Gamma_{2}$. If $\Gamma$ is the join of $\Gamma_{1}$ and $\Gamma_{2}$, then $A_{\Gamma} = A_{\Gamma_{1}} \times A_{\Gamma_{2}}$ and $X_{\Gamma} = X_{\Gamma_{1}} \times X_{\Gamma_{2}}$. We prove \setlength{\unitlength}{1cm} \begin{picture}(11,4.2) \thicklines \put(1.5,0.5){\large{$\Gamma$}} \put(0.5,2){\circle*{.1}} \put(2.5,2){\circle*{.1}} \put(1.5,3){\circle*{.1}} \put(0.5,2){\line(1,0){2}} \put(5.5,0.5){\large{$\Gamma'$}} \put(4.5,1.5){\circle*{.1}} \put(6.5,1.5){\circle*{.1}} \put(5.5,2.0){\circle*{.1}} \put(4.5,1.5){\line(1,0){2}} \put(4.5,3.5){\circle*{.1}} \put(6.5,3.5){\circle*{.1}} \put(5.5,4.0){\circle*{.1}} \put(4.5,3.5){\line(1,0){2}} \put(4.5,1.5){\line(1,1){2}} \put(6.5,1.5){\line(-1,1){0.92}} \put(4.5,3.5){\line(1,-1){0.92}} \put(9.5,0.5){\large{$\Gamma''$}} \put(8.5,1.5){\circle*{.1}} \put(10.5,1.5){\circle*{.1}} \put(9.5,2.0){\circle*{.1}} \put(8.5,1.5){\line(1,0){2}} \put(8.5,1.5){\line(2,1){1}} \put(10.5,1.5){\line(-2,1){1}} \put(8.5,3.5){\circle*{.1}} \put(10.5,3.5){\circle*{.1}} \put(9.5,4.0){\circle*{.1}} \put(8.5,3.5){\line(1,0){2}} \put(9.5,2.0){\line(2,3){0.43}} \put(9.5,2.0){\line(-2,3){0.43}} \put(10.5,3.5){\line(-2,-3){0.43}} \put(8.5,3.5){\line(2,-3){0.43}} \put(8.5,1.5){\line(2,5){0.77}} \put(9.5,4.0){\line(-2,-5){0.18}} \put(10.5,1.5){\line(-2,5){0.77}} \put(9.5,4.0){\line(2,-5){0.18}} \qbezier(8.5,1.5)(12,0.25)(10.5,3.5) \qbezier(9.3,1.15)(7,0.75)(8.5,3.5) \qbezier(10.5,1.5)(9.7,1.25)(9.7,1.25) \end{picture} \begin{proposition}\label{art} If $\Gamma$ is not a join, then $\mathrm{rank}(A_{\Gamma})=1$. \end{proposition} \begin{remark} If $\Gamma$ is the join of $\Gamma_{1}$ and $\Gamma_{2}$, then $$\mathrm{rank}(A_{\Gamma}) = \mathrm{rank}(A_{\Gamma_{1}}) + \mathrm{rank}(A_{\Gamma_{2}}) \geq 2.$$ \end{remark} The proof of Proposition \ref{art} is a direct consequence of a theorem due to Davis and Januszkiewicz(\cite{DavJan}). For a given graph $\Gamma$, we define two graphs $\Gamma '$ and $\Gamma''$ as follows : The vertex set of $\Gamma ''$ is $I \times \{0,1\}$, where $I$ is the vertex set of $\Gamma$. Two vertices $(i,1)$ and $(j,1)$ in $I \times 1$ are connected by an edge in $\Gamma''$ if and only if the corresponding vertices $i$ and $j$ span an edge in $\Gamma$. Any two distinct vertices in $I \times 0$ are connected by an edge. Finally, vertices $(i,0)$ and $(j,1)$ are connected by an edge if and only if $i \neq j$. The vertex set of $\Gamma '$ is $I \times \{-1, 1\}$. The subsets $I\times (-1)$ and $I \times 1$ both span copies of $\Gamma$. A vertex $(i,-1)$ in $I \times (-1)$ is connected to $(j,1)$ in $I \times 1$ if and only if $i \neq j$ and the vertices $i$ and $j$ span an edge of $\Gamma$. (See above for an example.) Then \begin{theorem}{\cite{DavJan}}\label{davjan} $A_{\Gamma}$ and $W_{\Gamma'}$ are subgroups of $W_{\Gamma''}$ of index $2^{I}$. \end{theorem} \begin{lemma}\label{finiteindex} $\Gamma$ is a join if and only if $\Gamma '$ is a join. \end{lemma} \begin{proof} For any subset $I' \subset I$, let $\Gamma_{I'}$ be a full subgraph of $\Gamma$ whose vertex set is $I'$. It is obvious that if $\Gamma$ is a join, then $\Gamma '$ is a join. Namely, if $\Gamma$ is a join of $\Gamma_{I_{1}}$ and $\Gamma_{I_{2}}$, then $\Gamma ' $ is a join of $\Gamma^{'}_{I_{1}\times\{-1,1\}}$ and $\Gamma^{'}_{I_{2}\times\{-1,1\}}$. Conversely, suppose $\Gamma'$ is a join of $\Gamma^{'}_{I'_{1}}$ and $\Gamma^{'}_{I'_{2}}$. Note that a vertex $(i,1) \in \Gamma^{'}_{I'_{1}}$ if and only if $(i,-1) \in \Gamma^{'}_{I'_{1}}$. Therefore, $\Gamma'$ is a join of $\Gamma^{'}_{I_{1}\times\{-1,1\}}$ and $\Gamma^{'}_{I_{2}\times\{-1,1\}}$ for some $I_{1}, I_{2} \subset I$ and $\Gamma$ is a join of $\Gamma_{I_1}$ and $\Gamma_{I_2}$. \end{proof} \begin{proof}[\,\,of Propositon \ref{art}] Suppose that $\Gamma$ is not a join. By Lemma \ref{finiteindex}, the corresponding graph $\Gamma'$ is not a join. It follows that $W_{\Gamma '}$ is irreducible. If $W_{\Gamma'}$ has a finite index free abelian subgroup $K$, then $K \cap A_{\Gamma}$ is also a finite index free abelian subgroup of $A_{\Gamma}$. But this is impossible: Since $\Gamma$ is assumed to be not a join, there are two vertices in $\Gamma$ which are not joined by an edge and they generate a non-abelian free subgroup of $A_{\Gamma}$. Call $a$ and $b$ for the generators. On the other hand, since $K \cap A_{\Gamma}$ is of finite index in $A_{\Gamma}$, there exist $N_1$ and $N_2$ such that $a^{N_1}, b^{N_2} \in K \cap A_{\Gamma}$ and $a^{N_1} b^{N_2} =b^{N_2}a^{N_1}$. This contradicts that $a$ and $b$ generate a free group. By Proposition \ref{cox}, $\mathrm{rank}(W_{\Gamma'})=1$, and hence, $\mathrm{rank}(A_{\Gamma})=1$. \end{proof} \begin{corollary} If $\Gamma$ is not a join, then $A_{\Gamma}$ is not commensurable (or quasi-isometric) to any uniform lattice in a non-compact connected semi-simple Lie group of higher rank. \end{corollary} \section{Relatively Hyperbolic Groups}\label{relhypgppre} Relatively hyperbolic groups are a generalization of hyperbolic groups. They were introduced by Gromov (\cite{Gro}) and many equivalent definitions have been developed by different authors in different contexts. See, for example, Farb(\cite{Far}), Bowditch(\cite{Bow}), Yaman(\cite{Yam}), Dru\c{t}u-Osin-Sapir(\cite{DruSap}), Osin(\cite{Osi}), Dru\c{t}u(\cite{Dru}), and Mineyev-Yaman(\cite{MinYam}). In this paper, we shall use the Bowditch's definition via geometrically finite convergence groups. Following \cite[Section 2]{Xie}, we recall the notion of relatively hyperbolic groups and the existence of an invariant collection of disjoint horoballs. See \cite{Bow} for details. Suppose that $M$ is a compact metrizable topological space. Suppose that a group $G$ acts by homeomorphisms on $M$. By definition, $G$ is $\emph{a convergence group}$ if the induced action on the space of distinct triples is properly discontinuous. In such a case, we call an element $g \in G$ a \emph{hyperbolic element} if it has infinite order and fixes exactly two points in $M$. A subgroup $H$ of $G$ is \emph{parabolic} if $H$ is infinite, fixes some point in $M$, and contains no hyperbolic elements. In this case, the fixed point of $H$ is unique. We call the point a \emph{parabolic point} and the nontrivial element in a parabolic subgroup a \emph{parabolic element}. It is necessary that the stabilizer of a parabolic point $\zeta$, $\mathrm{Stab}(\zeta)$, is a parabolic subgroup. A parabolic point $\zeta$ is \emph{a bounded parabolic point} if $\mathrm{Stab}(\zeta)$ acts properly and cocompactly on $M\setminus\{\zeta\}$. A point $\xi \in M$ is \emph{a conical limit point} if there exists a sequence $\{g_{n}\}$ in $G$ and two distinct points $\zeta, \eta \in M$, such that $g_{n}(\xi) \to \zeta$ and $g_{n}(\xi') \to \eta$ for all $\xi' \neq \xi$. Finally, a convergence group $G$ on $M$ is $\emph{a geometrically finite group}$ if each point of $M$ is either a conical limit point or a bounded parabolic point. \begin{definition}\label{relhypgp} A group $G$ is \emph{hyperbolic relative to a family of infinite finitely generated subgroups $\mathcal{G}$} if it acts properly discontinuously by isometries on a proper geodesic hyperbolic space $X$ such that the induced action on $\partial X$ is of convergence, geometrically finite, and such that the maximal parabolic subgroups are exactly the elements of $\mathcal{G}$. Elements of $\mathcal{G}$ are called \emph{peripheral subgroups}. \end{definition} It is known that all the definitions mentioned above are equivalent, provided that the group $G$ and all peripheral subgroups are infinite and finitely generated. But some authors do not assume that peripheral subgroups are infinite and finitely generated. In fact, it has been shown in \cite{Yam} that the finite generation of peripheral subgroups can be dispensed with. Also some definitions allow the elements of $\mathcal{G}$ to be finite. But, in \cite{Osi}, Osin proved that one can make $\mathcal{G}$ smaller so that all peripheral subgroups are infinite (or possibly empty). The followings are well-known examples of relatively hyperbolic groups. \begin{example} \begin{itemize} \item Hyperbolic groups: These are hyperbolic relative to $\mathcal{G} = \emptyset$. \item Geometrically finite isometry groups of Hadamard manifolds of negatively pinched sectional curvature: These are hyperbolic relative to the maximal parabolic subgroups. \item Free products of finitely many finitely generated groups: These are hyperbolic relative to the factors, since the action on the Bass-Serre tree satisfies the second definition of Bowditch. See \cite[Definition 2]{Bow}. \item Groups $G$ acting geometrically on a CAT(0) space $X$ which has the isolated flats property: In this case, $X$ is an asymptotically tree-graded space and $G$ is hyperbolic relative to the collection of virtually abelian subgroups of rank at least two. (See \cite{HruKle}) \end{itemize} \end{example} In the next subsection, we will prove that if $G$ is a relatively hyperbolic group with $|\mathcal{G}| \geq 2$, and at least one peripheral subgroup contains an element of infinite order, then $\mathrm{rank}(G) \leq 1$. The following theorem of Bowditch on the existence of an invariant collection of disjoint horoballs provides the crucial tool in our proof. Let $X$ be a $\delta$-hyperbolic geodesic metric space for some $\delta > 0$ and $\xi \in \partial X$. A function $h : X \to \mathbb{R}$ is \emph{a horofunction about $\xi$} if there exist constants $c_{1} = c_{1}(\delta), c_{2}=c_{2}(\delta)$ such that if $x,a \in X$ and $d(a,x\xi) \leq c_{1}$, for some geodesic ray $x\xi$ from $x$ to $\xi$, then $|h(a)-h(x)-d(x,a)| \leq c_{2}$. A closed set $B \subset X$ is \emph{a horoball about $\xi$} if there is a horofunction $h$ about $\xi$ and a constant $c=c(\delta)$ such that $h(x) \geq -c$ for all $x\in B$, and $h(x) \leq c$ for all $x \in X\setminus B$. In this case $\xi$ is called the center of the horoball and is uniquely determined by $B$. \begin{proposition}{\cite[Proposition 6.13]{Bow}}\label{horo} Let $G$ be a relatively hyperbolic group and $X$ a space on which $G$ acts as in Definition \ref{relhypgp}. Let $\Pi$ be the set of all bounded parabolic points in $\partial X$. Then $\Pi/G$ is finite. Moreover, for any $r>0$, there is a collection of horoballs $\mathcal{B}=\{B_{\xi} | \xi \in \Pi\}$ indexed by $\Pi$ with the following properties \begin{enumerate} \item $\mathcal{B}$ is $r$-separated, that is, $d(B_{\xi},B_{\eta}) \geq r$ for all $\xi \neq \eta \in \Pi$. \item $\mathcal{B}$ is $G$-invariant, that is, $g(B_{\xi}) = B_{g(\zeta)}$ for all $g \in G$ and $\xi \in \Pi$. \item $Y(\mathcal{B})/G$ is compact, where $Y(\mathcal{B}) = X \setminus \bigcup_{\xi \in \Pi} int(B_{\zeta})$. \end{enumerate} \end{proposition} Note that the intersection of any two peripheral subgroups is finite and there are finitely many conjugacy classes of peripheral subgroups. \subsection{Algebraic Rank of Relatively Hyperbolic Groups} In order to prove that a group has an algebraic rank $\leq 1$, we need to figure out the set $\mathcal{A}_{1}(G)$ and show that the group can be covered by finitely many translates of $\mathcal{A}_{1}(G)$. Also the procedure needs to be repeated for all finite index subgroups. We introduce two lemmas which enable us to find the elements such that the group can be covered by translates of $\mathcal{A}_{1}(G)$ by those elements. For a finite set of isometries $F$ of a metric space $X$ and $x \in X$, let $\lambda(x,F) = \mathrm{max}\{d(f(x),x)|f\in F\}$. \begin{lemma}{\cite{Kou}}\label{koubi} Let $X$ be a $\delta$-hyperbolic geodesic metric space and $G$ a group of isometries of $X$ with a finite generating set $S$. If $\lambda(x,S) > 100\delta$ for all $x \in X$, then $G$ contains a hyperbolic element $g$ such that $d_{S}(id,g) =1$ or $2$. \end{lemma} Hereafter, suppose that $G$ is a relatively hyperbolic group with $|\mathcal{G}| \geq 2$ and $X$ a proper $\delta$-hyperbolic geodesic space on which $G$ acts as in Definition \ref{relhypgp}. Also we assume that there is a peripheral subgroup in $\mathcal{G}$ containing elements of infinite order. Note that the existence of such a peripheral subgroup implies that there are two or more such subgroups by conjugation by hyperbolic elements. Lemma \ref{horo} implies that there is a $200\delta$-separated invariant collection of horoballs $\mathcal{B}$ centered at the parabolic points such that $Y(\mathcal{B})/G$ is compact. \begin{lemma}{\cite[Lemma 3.1]{Xie}}\label{xie} There exists a positive integer $k_1$ with the following property : for any infinite order element $\gamma \in G$ and any $x \in Y(\mathcal{B})$, there is some $k, 1 \leq k \leq k_{1}$, such that $d(\gamma^{k}(x),x) \geq 200\delta$. \end{lemma} Let $\mathcal{H}$ be the set of hyperbolic elements. \begin{proposition}\label{hypelerel} $\mathcal{H} \subset \mathcal{A}_{1}(G)$. \end{proposition} \begin{proof} Let $g \in \mathcal{H}$ be given and $A$ the two fixed points of $\langle g \rangle$ in $\partial X$. If $h \in G$ commutes with $g$, then $h$ fixes $A$ (See \cite[Corollary 2O]{Tuk}). Combining this with the fact that $\langle g \rangle$ is of finite index in the stabilizer $H=\{q \in G | qA=A\}$ (see \cite[Theorem 2I]{Tuk}), the centralizer of $g$ in $G$, $C_{G}(g)$ has a free abelian group of rank at most one as a finite index subgroup. Therefore, $\mathcal{H} \subset \mathcal{A}_{1}(G)$. \end{proof} Choose two elements of infinite order from two different peripheral subgroups and denote them by $h_{1}$ and $h_{2}$. We also denote the horoball stabilized by $h_{i}$ by $B_{i}, i=1,2$. Since $h_{i}$ is chosen to be of infinite order, $B_{i}$ is the only horoball stabilized by $h_{i}, i=1,2$. Then \begin{proposition}\label{relmain} Let $g \in G \setminus \mathcal{H}$ be an infinite order parabolic element. Then $h_{i}^{k}g$ is hyperbolic for some $i=1,2$ and for some $k, 1 \leq k \leq k_{1}$, where $k_{1}$ is the constant appearing in Lemma \ref{xie}. \end{proposition} \begin{proof} Without loss of generality, $h_1$ and $g$ are contained in different peripheral subgroups. Following Lemma \ref{xie}, consider $K=\langle h_{1}^{k}, g \rangle$. We prove that $\lambda(x,\{h_{1}^{k},g\}) > 100 \delta$ for any $x \in X$. Suppose that there exists some $x \in B$ for some $B \in \mathcal{B}$ such that $\lambda(x,\{h_{1}^{k},g\}) \leq 100\delta$. Then $h_{1}^{k}(B)=B$ and $g(B)=B$ (Note that $\mathcal{B}$ is $200\delta$-separated). It follows that the center of $B$ is fixed by $K$. Since $h_{1}^{k}$ and $g$ fix two different centers, this is a contradiction. By the choice of $k$, $d(h_{1}^{k}(x),x) \geq 200\delta$. Therefore, $\lambda(x, \{h_{1}^{k},g\}) \geq 200\delta$. Lemma \ref{koubi} implies that $h_{1}^{k}g$ or $gh_{1}^{k}$ is hyperbolic. (Note that both $h_{1}^{k}$ and $g$ are parabolic.) Furthermore, if $gh_{1}^{k}$ is hyperbolic, so is its conjugate $g^{-1}(gh_1^{k})g = h_{1}^{k}g$. In fact, suppose that $gh_{1}^{k}$ is hyperbolic and fixes exactly two distinct points $\alpha$ and $\beta$, then $h_{1}^{k}g$ fixes $h_{1}^{k}(\alpha)$ and $h_{1}^{k}(\beta)$. Conversely, suppose $h_{1}^{k}g$ fixes a point $\gamma \neq h_{1}^{k}(\alpha), h_{1}^{k}(\beta)$. Choose $\gamma'$ such that $h_{1}^{k}(\gamma')=\gamma$. Note that $\gamma' \neq \alpha, \beta$. But $h_{1}^{k}gh_{1}^{k}(\gamma') = h_{1}^{k}(\gamma') \Rightarrow gh_{1}^{k}(\gamma')=\gamma'$. Therefore, $\gamma'=\alpha$ or $\beta$, which is a contradiction. \end{proof} \begin{theorem}\label{relmain2} Suppose that $G$ is hyperbolic relative to a family $\mathcal{G}$ of infinite finitely generated subgroups. If $|\mathcal{G}| \geq 2$ and at least one subgroup in $\mathcal{G}$ contains an element of infinite order, then $\mathrm{rank}(G) \leq 1$ \end{theorem} \begin{proof} We decompose the set of torsion elements into $\mathcal{E} \coprod \mathcal{E'}$ as follows. A torsion element $g \in \mathcal{E}$ if and only if $g$ stabilizes both $B_{1}$ and $B_{2}$. Otherwise $g \in \mathcal{E}'$. Suppose that $g \in \mathcal{E}'$. Without loss of generality, assume that $g$ does not stabilize $B_{1}$. Then we have $\lambda(x,\{h_{1}^{k},g\}) \geq 100\delta$. In particular, $d(g(x),x) \geq 100\delta$ for $x \in B_{1}$. The same argument as in Proposition \ref{relmain} implies that $h_{1}^{k}g$ is hyperbolic. Since the intersection of two distinct peripheral subgroups is at most finite, $|\mathcal{E}| < \infty$, say $\mathcal{E} =\{l_{1}, \cdots, l_{n}\}$. Choose any hyperbolic element $h \in G$ and let $t_{i}=l_{i}h^{-1}$ for $i=1,\cdots, n$. By combining with Proposition \ref{relmain}, $$G = \mathcal{H} \bigcup (\bigcup_{i=1}^{k_1} h_{1}^{-i}\mathcal{H}) \bigcup (\bigcup_{i=1}^{k_1} h_{2}^{-i}\mathcal{H}) \bigcup(\bigcup_{i=1}^{n} t_{i}\mathcal{H}).$$ Therefore, $\mathrm{r}(G) \leq 1$. Next we need to prove that $r(T) \leq 1$ for any finite index subgroup $T$ in $G$. By taking the normal core of $T$, it suffices to show that $r(T) \leq 1$ for any finite index normal subgroup $T$ in $G$. Recall $r(G') \geq r(G)$ if $G'$ is a finite index subgroup of $G$. Let $T$ be a finite index normal subgroup in $G$ and $m=[G:T]$. Also let $\mathcal{H}_{T} = \mathcal{H} \bigcap T$. Recall that $\mathcal{H}$ is the set of hyperbolic elements in $G$. Then $\mathcal{H}_{T} \subset \mathcal{A}_{1}(T)$. It can be easily verified that all arguments in proving $r(G) \leq 1$ apply without any change to prove $r(T) \leq 1$, namely, \begin{itemize} \item $h_{i}^{m} \in T$ is an infinite order parabolic element and stabilizes a horoball $B_{i}$ for $i=1,2$. \item For any $g \in T\setminus \mathcal{H}_{T}$ of infinite order, $(h_{i}^{m})^{k} g \in \mathcal{H}_{T}$ for some $i=1,2$. \item One can decompose the set of torsion elements in $T$ as follows : $g \in \mathcal{E}_{T}$ if and only if $g$ stabilizes both $B_{1}$ and $B_{2}$. Otherwise $g \in \mathcal{E'}_{T}$. For any element $g$ in $\mathcal{E'}_{T}$, $(h_{i}^{m})^{k} g \in \mathcal{H}_{T}$ for some $i=1,2$. Since $\mathcal{E}_{T}$ is finite, one can choose any hyperbolic element in $T$ such that $\mathcal{E'}_{T}$ can be covered by finitely many translates of $\mathcal{H}_{T}$. \end{itemize} \end{proof} \subsection{$\mathrm{CAT}(0)$ Spaces with Isolated Flats} $\mathrm{CAT}(0)$ spaces with isolated flats were first introduced by Kapovich-Leeb and Wise, independently. In \cite{KapLee}, Kapovich and Leeb study a class of $\mathrm{CAT}(0)$ spaces in which the maximal flats are disjoint and separated by regions of strictly negative curvature. Since then, they have been studied by a number of authors, in particular, because of their strong connections to relatively hyperbolic groups. Throughout this subsection, a \emph{$k$-flat} is an isometrically embedded copy of Euclidean space $\mathbb{E}^{k}$ for $k \geq 2$. In particular, we don't consider a geodesic line as a flat. Let $\mathrm{Flat}(X)$ be the space of all flats in $X$ with the topology of uniform convergences on bounded sets. A $\mathrm{CAT}(0)$ space $X$ with a geometric group action has \emph{isolated flats} if it contains an equivariant collection $\mathcal{F}$ of flats such that $\mathcal{F}$ is closed and isolated in $\mathrm{Flat}(X)$, and each flat $F \subset X$ in the space is contained in a uniformly bounded tubular neighborhood of some $F' \in \mathcal{F}$. See \cite[Theorem 1.2.3]{HruKle} for equivalent formulations of $\mathrm{CAT}(0)$ spaces with isolated flats. Let $X$ be a CAT(0) space with isolated flats and $G$ be a group acting geometrically on $X$. One of main results in \cite{HruKle} is \begin{theorem}\cite[Theorem 1.2.1]{HruKle}\label{hrukle} The following are equivalent. \begin{enumerate} \item $X$ has isolated flats. \item $X$ is a relatively hyperbolic space with respect to a family of flats in $\mathcal{F}$. \item $G$ is a relatively hyperbolic group with respect to a collection of virtually abelian subgroups of rank at least two. \end{enumerate} \end{theorem} \begin{remark} \begin{itemize} \item In the second statement above, the term ``relatively hyperbolic" for metric spaces was introduced by Dru\c{t}u and Sapir. In \cite{DruSap}, they used the term ``asymptotically tree graded" for such spaces and proved that the metric and group theoretic notions of being relatively hyperbolic are equivalent for a finitely generated group with a word metric. \item If $X$ has isolated flats with respect to $\mathcal{F}$, then $\mathcal{F}$ is locally finite. Combining this with the Bieberbach Theorem shows that each flat in $\mathcal{F}$ is $G$-periodic with virtually abelian stabilizer. Note that the geometric action of $G$ on $X$ induces a quasi-isometry and being relatively hyperbolic with respect to quasiflats is a geometric property. (See \cite[Theorem 5.1]{DruSap}.) This quasi-isometry takes $\mathcal{F}$ to the left cosets of a collection of virtually abelian subgroups of rank at least two. See \cite[Section 3, 4]{HruKle} for details. \end{itemize} \end{remark} \begin{proposition}\label{catmain} Let $G$ be a group acting geometrically on a $\mathrm{CAT}(0)$ space with isolated flats and $|\mathcal{F}| \geq 2$. Then ${\mathrm{rank}}(G) \leq 1$. \end{proposition} \begin{proof} By Theorem \ref{hrukle}, $G$ is hyperbolic relative to a collection of virtually abelian subgroups of rank at least two. Since we assume that $|\mathcal{F}| \geq 2$, there are at least two peripheral subgroups in $G$. Proposition \ref{relmain2} applies that $\mathrm{rank}(G) \leq 1$. \end{proof} \begin{remark} In the case that $\mathcal{F}$ consists of a single flat $F$, one can conclude that $G$ acts geometrically on $F$. Therefore, $\mathrm{rank}(G)$ is equal to $dim(F) \geq 2$ by \cite{BalEbe}. \end{remark}
\section{Introduction} Coupling individual oscillators can lead to amplitude death, the cessation of oscillations due to stabilization of a stationary state\cite{Atay2010}. This phenomenon has been observed in chemical oscillators \cite{Bar-Eli1985}, electronic circuits \cite{Reddy2000}, thermo-optical oscillators \cite{Herrero2000}, and coupled lasers \cite{Prasad2003}. Amplitude death can be beneficial and is exploited for instance in feedback control applications \cite{Konishi2008}, but can also be detrimental and even lethal if it occurred in interacting cardiac cells \cite{Atay2006JDE}. In mathematical models, amplitude death is commonly studied by investigating the conditions under which an unstable steady state of an isolated system is stabilized by coupling. It was found that for amplitude death to occur, either the natural frequencies of the coupled systems need to be sufficiently disparate \cite{Yamaguchi1984,Shiino1989,Aronson1990,*Ermentrout1990,Mirollo1990} or the coupling needs to be time-delayed \cite{Reddy1998}. While amplitude death can be observed already in systems of two oscillators, richer behavior is observed when complex networks of oscillators are considered. In such networks the nodes represent individual oscillators, whereas the links represent coupling terms. A central question is then how the dynamics are affected by network topology, the specific structure of nodes and links. For a given network of $N$ oscillators this topology can be captured by the adjacency matrix $\mat{A}$, a $N \times N$ matrix with $A_{ij}=1$ if a link exists from node $j$ to node $i$ and $A_{ij}=0$ otherwise. A related matrix known to affect many network properties is the network Laplacian $\mat{L} = \mat{A} - {\rm \bf D}$. Here, $\mat{D}$ is a diagonal matrix, with $D_{ii} = k_i$, the degree of node $i$, i.e.~the number of nodes to which the focal node is connected. Delay induced amplitude death is often studied in systems of coupled limit-cycle oscillators, such as Stuart-Landau oscillators. Studies on globally connected networks \cite{Reddy1998,*Reddy1999} and rings \cite{Dodla2004} showed that amplitude death occurs inside islands in the parameter space of coupling strength and delay. While distributed delays seem to enlarge these islands \cite{Atay2003PRL}, gradient instead of diffusive coupling has been found to suppress amplitude death in rings \cite{Zou2010}. Further, transients between regimes of partial and complete amplitude death have been studied by numerical simulations \cite{Yang2007}. Also the effect of delayed self-feedback has been studied in a single and two delay-coupled Stuart-Landau oscillators \cite{Reddy2000PhysD, DHuys2010}. Delay-induced amplitude death has been explored in detail in discrete-time maps \cite{Atay2004,*Atay2006SIAM,Masoller2005,Gong2008,Ponce2009} and in systems of oscillators modeled by multi-dimensional ordinary differential equations, such as the R\"{o}ssler and Lorenz oscillators \cite{Atay2003PhysD,Atay2006JDE,Konishi2004,*Konishi2005,Prasad2005}. For a model of coupled maps, it was shown that the dynamics is governed by the largest eigenvalue of the network's Laplacian \cite{Atay2006SIAM}. A similar result was shown in time-continuous system \cite{Atay2006JDE}. Previous studies have shown impressively that analyzing amplitude death in simple systems, such as simple oscillators or maps can yield a deep understanding, whereas coupling more complex systems such as R\"{o}ssler or Lorenz oscillators points to additional complexities. A middle way is perhaps offered by using delay oscillators, such as Mackey-Glass \cite{Mackey1977,Farmer1982} and the Ikeda \cite{Ikeda1987} oscillators. Because even a single-variable delay-differential equation (DDE) constitutes an infinite dimensional dynamical system, a single DDEs can show sustained oscillations, quasi-periodicity, and chaos. Studying amplitude death in delay-coupled systems of such delay oscillators offers the opportunity to consider a network of relatively complex coupled system, while keeping the equations concise. Although a whole zoo of different synchronization types have been found for systems of coupled delay-oscillators without coupling delay \cite{Pyragas1998,Voss2000,Zhan2003,Li2004,Shahverdiev2005,Chen2007}, delay-coupled delay oscillators, have been less explored. One exception is \cite{Konishi2008} where a system of two mutually delay-coupled delay oscillators was investigated numerically, but more complex coupling topologies were not considered. A model for complex networks of general scalar delay-coupled delay oscillators was proposed in two previous publications of the present authors \cite{Hoefener2011,*Hoefener2012}. The latter paper highlighted that certain meso-scale structures in networks have a distinct impact on the network-level dynamics, whereas the former discussed the effect of the degree distribution, the probability distribution of node degrees. For degree-homogeneous networks (DHONs), where every node is connected to the same number of of other nodes, the dynamics of delay-coupled delay oscillators were studied analytically. This revealed the bifurcation lines at which the dynamics of the system change qualitatively, and linked them to certain eigenvalues of the networks adjacency matrix. For degree-heterogeneous networks (DHENs), numerical explorations showed qualitatively similar dynamics, with minor corrections. In the present paper, we use the previously proposed model to study amplitude death. For this purpose, we focus on a regime where the uncoupled systems show non-stationary dynamics. Then, we study the parameter space of coupling strength and coupling delays to find stable areas, which correspond to regions of amplitude death. For DHONs, we find that there exists a region of amplitude death for coupling delays larger than a certain threshold value, which approaches zero as $1/(dk)$ where $d$ is the common degree of nodes and the coupling strength $k$ approaches infinity. By using a numerical sampling method, we confirm this result for degree-heterogeneous networks, where $d$ is now the mean degree of the network. This shows that in the limit of large coupling strength the dynamics is governed by a global property of the network, while all other topological properties become negligible. \section{Model} We consider networks of $N$ nodes $i$, carrying a dynamical state $X_i$, which can represent for instance an ecological population or the abundance of RNA molecules in a genetic oscillator. The variable is subject to gain and loss terms, where the loss is instantaneous, while the gain term of $X_i$ at time $t$ depends on the value of $X_i$ at time $t-\tau$. Furthermore, $X_i$ continuously diffuses to topological neighbors causing an additional loss. The exported material arrives at the respective neighboring node $j$ after a travel time delay $\delta$. Thus, the dynamics is governed by \begin{equation}\label{eq:DDE} \dot{X}_i=G(X_i^{\tau})-L(X_i)+\sum_j \left(A_{i j} F(X_j^{\delta})-A_{j i}F(X_i)\right), \end{equation} where $G$, $L$ and $F$ are positive functions describing growth, loss, and coupling, respectively. Instead of restricting these functions to specific functional forms, we consider a general model comprising the whole class of models that includes well-studied examples such as the delay-coupled Mackey-Glass and the Ikeda systems. Here we only consider networks for which the number of outgoing links $d_i^\mathrm{out}$ equals the number of incoming links $d_i^\mathrm{in}$ for each node $i$, such that $d_i^\mathrm{out}=d_i^\mathrm{out}=:d_i$, where $d_i$ is the degree of node $i$. We note that this class includes among all bidirectionally coupled networks. For all networks in the class considered, a homogeneous steady state $X_i*=X^*$ exists, for each steady state $X^*$ of the isolated system, which satisfies $G(X^*)=L(X^*)$. We apply the method of generalized modeling \cite{GMPRE,GMSCIENCE}, which analyzes the dynamics of general system by a direct parametrization of the Jacobian matrix ${\rm \bf J}$, with $J_{ij} = \partial \dot{X}_i / \partial X_j$ in $X^*$. This matrix constitutes a local linearisation and thus governs the dynamics of the system close to a steady state under consideration \cite{Kuznetsov}. In order to express the Jacobian in terms of interpretable parameters, we normalize the system to an arbitrary homogeneous steady state ${X_i}^*$ and introduce normalized variables $x_i=X_i/X^*$ and normalized functions $f(x_i)=F(x_i X^*)/F(X^*)$, denoted by lower case symbols. Using $X^*=X^{\tau *}=X^{\delta *}$, we rewrite Eq.~\eqref{eq:DDE} as \begin{equation}\label{eq:DDE_norm} \dot{x}_i=\alpha(g(x_i^{\tau})-l(x_i))-d_i \beta f(x_i)+\beta\sum_j A_{i j} f(x_j^{\delta}). \end{equation} The quantities $\alpha=G(X^*)/X^*=L(X^*)/X^*$ and $\beta=F(X^*)/X^*$ are unknown constants and can thus be interpreted as parameters of the Jacobian. These parameters have the unit of inverse time and describe turnover rates. In the following we set $\alpha=1$ by a time scale normalization. In order to analyze the stability of the homogeneous steady state, we proceed as in \cite{MacDonald1989} and assume that, close to the steady state, the DDE has exponential solutions $\vec{y}(t)=\vec{y}(0)\exp(\lambda t)$, where $\vec{y}(t)=\vec{x}(t)-1$ is a small perturbation from the steady state at $\vec{x}=\vec{1}$. Using this ansatz together with the linearization of Eq.~\eqref{eq:DDE_norm} the Jacobian matrix has the entries \begin{equation} \label{eq:Jac} \begin{array}{r c c l} J_i^\mathrm{d}&:=&\gp\exp(-\lambda\tau)-\ensuremath{l'}-d_i\beta\ensuremath{f'} & \text{for } i=j\\ J^\mathrm{o}&:=&\beta\ensuremath{f'}\exp(-\lambda\delta) & \text{for } A_{i j}=1\\ 0 & & & \text{otherwise} \end{array} \end{equation} where $\lambda$ is an eigenvalue of the Jacobian and thus a solution of the characteristic polynomial $P(\lambda)=\det(\lambda\mat{I}-\mat{J})$. The quantities $\gp,\ensuremath{l'},\ensuremath{f'}$ denote derivatives of the normalized functions in the steady state, or equivalently logarithmic derivatives of the original functions, i.e.~$\gp=\left.\partial g / \partial x \right|_{1}=\left.\partial \log{(G)} / \partial \log (X)\right|_{*}$. Although we do not specify the functions or the steady state under consideration, these derivatives are formally constants and can thus be interpreted as unknown parameters of the system. Closer inspection reveals that these so-called elasticities are generally easily interpretable in the context of applications and have a number of additional benefits\cite{GMPRE}. Since $\ensuremath{f'}$ only appears in products together with $\beta$, we introduce the effective coupling strength $k:=\beta\ensuremath{f'}$. The steady state under consideration is asymptotically stable if all roots of the characteristic polynomial have negative real-parts and it is unstable if at least one root has a positive real part \cite{Atay2010}. In DDEs the explicit appearance of $\lambda$ in the Jacobian turns the characteristic polynomial into a transcendental function, which may have an infinite number of solutions. To make progress analytically it is therefore advantageous to directly compute the bifurcation points in parameters space at which eigenvalues acquire positive real parts. This is general simpler than computing the spectrum for a single parameter set. In numerical explorations the computation of the spectrum can be avoided by checking for positive eigenvalues with a method based on Cauchy's argument principle \cite{Luzyanina1996,Hoefener2011}. \section{Bifurcations of degree homogeneous networks} In a previous publication \cite{Hoefener2011}, we demonstrated that the characteristic equation $P(\lambda)=0$ of DHONs can be decomposed into $N$ independent equations, each of which corresponds to a single eigenvalue of the adjacency matrix. In this section we briefly recapitulate this derivation that will form the basis for our discussion of amplitude death below. Because the treatment involves both the eigenvalues $\lambda_i$ of the system's Jacobian matrix, and the eigenvalues $c_i$ of the system's adjacency matrix, we refer to $\lambda_i$ as dynamic eigenvalues and to $c_i$ as topologial eigenvalues. In a DHON all diagonal elements $J_i^\mathrm{d}$ are identical. This allows us to substitute $\lambda-J^\mathrm{d}$ in the characteristic polynomial $P(\lambda)$ by $z$, which yields $P'(z)=\det(z\mat{I}-J^\mathrm{o}\mat{A})$. The roots of $P'(z)$ are given by $c_i J^\mathrm{o}$, where $c_i$ denotes one of the $N$ topological eigenvalues. The back-substitution yields $N$ independent equations, \begin{equation}\label{eq:EV_DHON} \lambda=J^\mathrm{d}(\lambda)+c_i J^\mathrm{o}(\lambda). \end{equation} We note that the topological eigenvalues $c$ are in general complex, so that $c=|c|\expe{i\ensuremath{\psi^\mathrm{c}}}$, where $\ensuremath{\psi^\mathrm{c}}$ is the complex phase of the topological eigenvalue. For determining the stability of the system, we recall that the stability changes at bifurcation points where a dynamical eigenvalue crosses the imaginary axis and becomes purely imaginary, so that $\lambda=i \omega$. By separating the real and imaginary part of Eq.~\eqref{eq:EV_DHON} we obtain \begin{eqnarray} \label{eq:Re}0&=&\gp\cos(\phi)-\ensuremath{l'}-d\ensuremath{k}+|c|\ensuremath{k}\cos(\psi),\\ \label{eq:Im}\omega&=&-\gp\sin(\phi)-|c|\ensuremath{k}\sin(\psi), \end{eqnarray} with $\phi=\omega\tau$ and $\psi:=\omega\delta-\ensuremath{\psi^\mathrm{c}}$. The equation system contains three unknown variables, $\phi$, $\psi$ and $\omega$. Given a solution triplet $(\phi,\psi,\omega)$, we can find another solution $(-\phi,-\psi,-\omega)$. The solutions with negative $\omega$ trace the same bifurcation lines as those with positive $\omega$. We can thus restrict the analysis to solutions $\omega\geq 0$ without missing bifurcations. Further, we are only interested in solutions with $\tau>0$, and therefore $\phi>0$. By choosing $\phi$ as a free parameter, we can find a parametric representations of the bifurcation lines, but need to distinguish three cases depending on the value of the topological eigenvalue $c$. For the case $c=0$ the Eqs.~(\ref{eq:Re},\ref{eq:Im}) are independent of $\psi$ and $\delta$, respectively. Thus, the bifurcations are vertical lines in the $(\ensuremath{k},\delta)$-plane, with \begin{equation}\label{eq:beta_c0} \ensuremath{k}=\frac{1}{\tau}\frac{h}{d}, \end{equation} where $h:=(\gp\cos(\phi)-\ensuremath{l'})\tau$ and $\phi$ has to satisfy $f(\phi):=\phi+\gp\tau\sin(\phi)=0$. The equation $f(\phi)=0$ may generally have several solutions. However, for the parameters used throughout this paper the solution is unique. For the case $|c|=d$, we find \begin{equation}\label{eq:beta_cd} \ensuremath{k}=\frac{1}{\tau}\frac{f^2+h^2}{2dh}. \end{equation} In order to obtain positive solutions, $h(\phi)$ needs to be positive. Therefore, we have to restrict $\phi$ to the intervals $I^\ensuremath{k}_r=[\phi^\ensuremath{k}_r,\overline{\phi^\ensuremath{k}_r}]$, with $\phi^\ensuremath{k}_r=2\pi r + \phi^\ensuremath{k}$, $\overline{\phi^\ensuremath{k}_r}=2\pi(r+1)-\phi^\ensuremath{k}$ and $\phi^\ensuremath{k}=\cos^{-1}(\ensuremath{l'}/\gp)$. The parameter $r$ that appears here is the first of two integer parameters that we introduce to enumerate the bifurcation lines. Finally, for the case $0<|c|<d$, we find \begin{equation}\label{eq:beta} \ensuremath{k}_\mathrm{1,2}=\frac{1}{\tau}\frac{dh\pm\sqrt{a}}{d^2-\abs{c}^2}, \end{equation} with \begin{equation} \label{eq:a}a=d^2h^2-(d^2-|c|^2)(f^2+h^2). \end{equation} For obtaining real and positive solutions, we require not only $h(\phi)>0$ but also $a(\phi)>0$. This imposes additional constraints on $\phi$ that can be calculated numerically. In contrast to the case $|c|=d$, valid solutions can only be found inside a finite number of intervals $I^\mathrm{\ensuremath{k}}_r$. We remark that the case $|c|>d$ does not need to be considered because the eigenvalues $c$ of the adjacency matrix cannot exceed the highest node degree, which is $d$ in DHONs. For the cases above, we compute the corresponding values of $\delta$ at which the bifurcation occurs from Eqs.~(\ref{eq:Re},\ref{eq:Im}). We find \begin{align} \label{eq:delta}\delta^\mathrm{L,R}&=\frac{\psi^\mathrm{L,R}+\psi_{\rm c}+2\pi s}{\phi}\tau,\\ \label{eq:psi}\psi^\mathrm{L,R}&=\pm\cos^{-1}\left(\frac{d}{|c|}-\frac{h}{|c|\ensuremath{k}}\right), \end{align} where $s$, is the second integer parameter introduced to enumerate solution branches. Furthermore, the indices L, respectively R, are introduced to denote the positive, respectively negative, sign branch of $\psi$. To obtain physical solutions we consider the L-branch for all $f(\phi)<0$ and the R-branch otherwise. Evaluating these, we find positive values of $\delta$ for non-negative integers $s$. One implication of Eq.~\eqref{eq:psi} is revealed when we consider that the arcus cosine needs to be smaller or equal to one. This is only possible if $\ensuremath{k}\le\ensuremath{k}^\mathrm{max}$ with \begin{equation} \ensuremath{k}^\mathrm{max}=-\frac{l+g}{d-|c|}. \end{equation} Thus, bifurcation lines corresponding to topological eigenvalues $c$ with $|c|<d$ can only be found for finite values of $\ensuremath{k}$. Hence, only eigenvalues with $|c|=d$ can affect the stability for sufficiently large coupling strength, independently of the coupling topology. \section{Amplitude death in degree homogeneous networks} Having derived the results in the previous section, we now turn to the analysis of amplitude death. In the parameter range $\gp<-\ensuremath{l'}<0$, we find that an isolated node is unstable if $\tau>\tau^*$, where $\tau^*=\tau^*_{r=0}$ and \begin{equation} \tau^*_r=\frac{\phi^\ensuremath{k}+2\pi r}{\sqrt{\gp^2-\ensuremath{l'}^2}}. \end{equation} The isolated system thus exhibits non-stationary (e.g.~oscillatory) dynamics when the reproductive delay $\tau$ is chosen sufficiently large. The stationary solutions can then potentially be stabilized by coupling the oscillators. In general stabilization will depend on the coupling topology, coupling delays $\delta$, and the coupling strength $k$. In the following we seek to identify the effect of the coupling topology on the areas in the $(\ensuremath{k},\delta)$-space where the homogeneous steady state is stable, such that amplitude death can occur. \begin{figure} \centering \includegraphics{biflines} \caption{Bifurcation diagrams indicating the dynamics for a fully connected network of 3 nodes (a) and a ring of four nodes (b). Lines mark bifurcation points corresponding to different topological eigenvalues $c_i$ (see text). Labels of the form $(r,s)$ on some of the lines state the corresponing indices $r$ and $s$ enumerating the analytical solution branches. For the limit of high coupling strength the bifurcation lines asymptotically approach limiting values, some of which are indicated on the right border of a) by symbols of the form $\delta_{r,s}^\mathrm{b,t}$, where $b$ and $t$ denote bottom and top branches, respectively. Gray dots mark stable states that were found numerically by uniformly sampling the ($k$,$\delta$) plane. The two lower plots (c,d) show the dependence of limiting values $\delta_{r,s}^\mathrm{b,t}$ on the ratio of local parameters $-l/g$ (c) and the dependence of $\delta_{r,r+1}^\mathrm{b}$, $\delta_{r,r}^\mathrm{t}$ on the index $r$. Parameters: $\gp=-1,\ensuremath{l'}=0,\tau=5$.} \label{fig:BifLines} \end{figure} Our analytical treatment above has identified a connection between the topological eigenvalues and the dynamical stability. For illustration of the analytical results we consider two specific topologies: A fully connected networks of 3 nodes and a ring of four nodes. Let us start with the fully connected network (Fig.~\ref{fig:BifLines}a). All such networks have a topological eigenvalue $c=d=N-1$ and a $d$-fold degenerate topological eigenvalue $c=-1$. For $c=-1$ computation of the bifurcation lines only yields physical solutions for the $r=0$ branch. In this branch different values of $s$ generate different segments of bifurcation lines, which connect such that a single long bifurcation line running from $\delta=0$ to $\delta=\infty$ is formed. Because $|c|<d$, all bifurcation points in this line occur at finite values of the effective coupling strength $\ensuremath{k}$. For the remaining eigenvalue $c=d$, the situation is more complex. Here, computation of the bifurcation points yields a family of separate tongues corresponding to different values of $r$ and $s$. Similar families of bifurcation lines have been observed in other delay systems \cite{SchollTwoDelays}. In the present paper, each of the tongues consists of two segments, to which we refer as the bottom and top part respectively. An exceptional solution is again the case $r=0$. Here, the lower boundary of the tongue is formed by the \emph{top} branch of the $(r=0,s=0)$ solution, while the upper boundary is formed by the \emph{bottom} branch of the $(0,1)$ solution. One can thus think of the $r=0$ case as an inside-out tongue, which provides the right intuition for understanding the results below. Let us now turn to the ring network (Fig.~\ref{fig:BifLines}(b)). All even rings are bipartite networks, which means that it is possible to color the nodes in two colors such that no node has a topological neighbor of the same color. Bipartite degree homogeneous networks have topological eigenvalues at $c=d$ and $c=-d$, additionally the 4-ring has a two-fold degenerate eigenvalue at $c=0$, which results from symmetry. For the topological eigenvalue $c=0$ only the ($r=0$,$s=0$) solution corresponds to a physical bifurcation line. This line occurs at a constant value of the coupling strength. The eigenvalues with $c=d$ and $c=-d$ each generate a family of tongues that are similar to those in the fully connected network. However, the tongues with $c=-d$ are shifted such that they are centered on the gaps between the tongues with $c=d$. From the corresponding eigenvectors one can see that the family of tongues with $c=d$ corresponds to an instability with respect to in-phase oscillations, whereas the tongues with $c=-d$ correspond to anti-phase oscillations. It is intuitive that this instability arises from the eigenvalue $c=-d$ which is directly linked to bipartiteness, because none-bipartite networks could not sustain anti-phase oscillations. The analytical results above can be confirmed by a numerical sampling procedure, in which we pick parameter sets uniformly from the $(k,\delta)$-plane and evaluate their stability numerically \cite{Hoefener2011}. For visualization we plot only those points that are found to be stable. The results shown in Fig.~\ref{fig:BifLines} shows that such stable parameter combinations are found only in certain regions that are sharply delineated by the theoretically predicted bifurcation lines. The figure shows that stability requires that the respective point lies inside all tongues with $r=0$ but outside all tongues with $r \neq 0$, which confirms the inside-out nature of the tongues with $r=0$ mentioned above. Thus for stable points, the bifurcation lines corresponding to $r=0$ impose a minimum coupling strength, whereas the solutions with $r\neq 0$ impose a maximum coupling strength. We note that the $c=0$ topological eigenvalue never causes a bifurcation of a stable solution. This eigenvalue gives rise to a physical branch with $r=0$ and thus potentially imposes a lower limit for the coupling strength in stable states. However, this limit is always below a higher limit imposed by the $r=0$ branches of other eigenvalues, and thus has no direct relevance for stability. This conforms to the general observation that only the largest positive and the smallest negative topological eigenvalue cause bifurcations that border stable regions. For DHONs it is probably possible to prove this rigorously, but the proof is beyond the scope of the current paper. Nevertheless, it is interesting to note that in all DHONs the largest positive eigenvalue is $c=d$, whereas the smallest negative eigenvalue is maximal for fully connected networks ($c=0$) and minimal for bipartite networks ($c=-d$). This illustrates why amplitude death is most likely in the fully connected network and least likely in bipartite networks. All other networks fall between these two extreme cases, which motivated our choice of examples. \section{Strong coupling limit} In the following, we focus on large coupling strength $\ensuremath{k}$. Above we showed already that only those bifurcation lines that correspond to topological eigenvalues with $|c|=d$ can extend to arbitrarily large coupling strength. Except in bipartite networks, this condition is only satisfied for the eigenvalue $c=d$. In the following we thus focus solely on this eigenvalue. In the limit $\ensuremath{k} \to \infty$, the bifurcation lines approach constant values of the coupling delay $\delta_{r,s}^\mathrm{b,t}$, respectively, where t and b denote values for top and bottom branches. For calculating $\delta_{r,s}^\mathrm{b,t}$, we study Eq.~\eqref{eq:beta_cd} and find that $\ensuremath{k}$ becomes infinitely large if $\phi$ approaches $\phi^\mathrm{\ensuremath{k}}_r$ and $\overline{\phi^\mathrm{\ensuremath{k}}_r}$. Using Eq.~\eqref{eq:delta} we obtain \begin{equation}\label{eq:delta_tb} \delta^\mathrm{t}_{r,s}=\frac{2\pi s +\psi^\mathrm{c}}{2\pi r +\phi^\ensuremath{k}}\tau,\qquad \delta^\mathrm{b}_{r,s}=\frac{2\pi s +\psi^\mathrm{c}}{2\pi (r+1) -\phi^\ensuremath{k}}\tau. \end{equation} We note that both equations scale linearly with $\tau$, such that all tongues scale with the internal delay of the oscillators. Furthermore, recalling that $\phi^\ensuremath{k}=\cos^{-1}(\ensuremath{l'}/\gp)$, we see that for $\gp=-\ensuremath{l'}$, the bottom and top limits are identical for each bifurcation line $(r,s)$, such that the unstable areas disappear (see also Fig.~\ref{fig:BifLines}(c) for selected lines). Let us now investigate the question whether in the limit of large coupling strength amplitude death is still possible. Already visual inspection of Fig.~\ref{fig:BifLines}(a) reveals that there are large channels of stability between the tongues, which seem to extend to high values of $\ensuremath{k}$. In particular, notable is the \emph{1:1-resonant channel} around $\delta=\tau$ and the bottom channel at small values of $\delta$. As we increase the coupling strength, all stable channels are successively narrowed down as new tongues, corresponding to branches with higher $r$, become relevant. For instance, the 1:1-resonant channel is bordered from below by the top branch of $(r,r)$ and from above by the bottom branch of $(r,r+1)$; the bottom channel is bordered from below by the top branch of $(0,0)$ and from above by the bottom branch of $(r,1)$. \begin{figure}[ht!] \centering \includegraphics{scaling} \caption{Amplitude death for large coupling strength. (a) Bifurcation lines from the eigenvalue $c=d$ (solid lines) and $c=-d$ (dashed lines). The approximate tip positions of tongues have been marked by circles ($c=d$) and squares ($c=-d$). The orange dashed line marks the analytical approximation to the bottom boundary of the stable region. For sufficiently large $k$, all degree homogeneous networks are stable between this bifurcation line and the tongues corresponding to $c=-d$. Non-bipartite networks are even stable up the tongues corresponding to $c=d$. (b) Numerical results from uniform sampling of the space shown for random networks with $N=10$ nodes and $K=15$ edges. Red dots mark unstable parameters points, such that the large white regions mark stable areas that correspond to amplitude death. In particular, for sufficiently $k$, no unstable networks are found for $\delta$ between the bottom bifurcation line with ($d=2K/N$) and the tip positions for $c=-d$. Most networks are stable up to the tip positions for $c=d$. In the double logarithmic plot the region of amplitude death growth with increasing $k$ as its boundaries approach power-laws with exponent $-1$ and $-1/2$ respectively, in agreement with analytical predictions. Parameters are $\gp=-1,\ensuremath{l'}=0,\tau=5$.}\label{fig:Scaling} \end{figure} We now focus specifically on the bottom channel, which comes arbitrarily close to the case of zero coupling delay and thus the well studied case of networks of delay oscillators without coupling delay. We start by considering the lower boundary of this channel by studying the asymptotic behavior of the $(0,0)$ bifurcation line. We note that in the relevant limit this is given by the L-branch of Eqs.~(\ref{eq:delta},\ref{eq:psi}) with $\phi$ approaching $\phi^\mathrm{\ensuremath{k}}$. In this case $h\to 0$, so that we can approximate Eq.~\eqref{eq:psi} by using $\cos^{-1}(1-x)\approx\sqrt{2|x|}$. With $\tau^*=\phi^\ensuremath{k}/\sqrt{\gp^2-\ensuremath{l'}^2}$ and $\phi^\ensuremath{k}=\cos^{-1}(\ensuremath{l'}/\gp)$, we find the critical coupling delay \begin{equation}\label{eq:delta_approx} \delta=\left(\frac{\tau}{\tau^*}-1\right)\frac{1}{d\ensuremath{k}}, \end{equation} where we dropped the indices for simplicity. This result shows that the critical $\delta$, which marks the lower boundary of the bottom channel, is inversely proportional to the effective coupling strength $d\ensuremath{k}$. For finding the critical $\delta$ that marks the upper boundary of the bottom channel, we approximate the tip positions of the tongues $(r,1)$ in the limit of large $r$. In this case we can approximate Eq.~\eqref{eq:beta_cd} with \begin{equation}\label{eq:beta_approx} k\approx\frac{\phi^2}{2dh\tau}. \end{equation} Using that $\phi$ is large, we find local minima at $\phi=(2r+1)\pi$. Inserting this result into the R-branch of Eqs.~(\ref{eq:delta},\ref{eq:psi}) yields an expression for the $\delta$-value. For large $\ensuremath{k}$, we see from Eq.~\eqref{eq:delta} that $\delta$ is inversely proportional to $\phi$, while we see from Eq.~(17) that $\ensuremath{k}$ is proportional to $\phi^2$. Thus, $\delta \propto 1/\sqrt{\ensuremath{k}}$. In summary, these results show that in the limit of large coupling strength there is a region of amplitude deaths at very small coupling delays. The onset of amplitude death occurs at a value of the coupling delay that scales as $1/k$, whereas the width of the amplitude death region (in terms of the coupling delay) scales as $1/\sqrt{k}$. We note that the reasoning presented in this section is not limited to the example topologies discussed above, but holds for all degree-homogeneous networks. While additional topological eigenvalues with $|c|=d$ could exist in directed or bipartite networks, the corresponding tongues would narrow the channels of amplitude death by a factor, but would show the same scaling behavior. \section{Amplitude death in heterogeneous network} As the final step of our analysis we investigate amplitude death in degree heterogeneous networks (DHENs). In particular we focus again on the limit of large coupling strength and small coupling delays that corresponds to the bottom channel computed analytically in DHONs. For DHENs comparable analytical calculations are not easily possible. We therefore explore the stability of these systems numerically. Figure Fig.~\ref{fig:Scaling} shows a scatter plot in which each point correpsonds to specific values of $\ensuremath{k}$ and $\delta$, sampled uniformly, and a random Erd\H{o}s-Reny\'{i} network with $N=10$ nodes and $K=15$ links. For each of these randomly generated sample networks we used the numerical method described in \cite{Luzyanina1996,Hoefener2011} to compute the stability. In the figure unstable samples where marked by a red dot, whereas stable regions remain white. From the scatter plot we see that for sufficiently large $\ensuremath{k}$, no unstable networks can be found in a large region at high coupling strength and small coupling delays. This region corresponds remarkably well to the analytical results that have been obtained for DHONs. In particular there is a sharp loss of stability at he bottom bifurcation line. For sufficiently high coupling strength almost all unstable samples fall into the tongues of instability generated by the eigenvalue $c=d$ in DHONs, whereas few fall into the tongues with $c=-d$. These samples can probably be explained by the random creation of bipartite networks, and would hence be virtually absent in larger networks. \begin{figure} \centering \includegraphics{avg_deg.pdf} \caption{Onset of amplitude death. Scatter plot of parameter values for which one randomly drawn network with $N=10$ nodes and $K$ links is stable and another one is unstable. Thus the dots mark the border between stable and unstable regions. The borders of network ensembles with $K=14$, $K=15$ and $K=16$ approach distinct lines, which are given by Eq.~\eqref{eq:delta_approx} with $d=2K/N$ (horizontal lines). The $\delta$-values are normalized by the line for $d=3$. Parameters are $\gp=-1,\ensuremath{l'}=0,\tau=5$.}\label{fig:AvgDeg} \end{figure} As a further test we study the onset of amplitude death as the bottom (0,0) bifurcation line in more detail. To capture the onset of amplitude death we are interested in those systems values where for given values of $\delta$ and $\ensuremath{k}$ the stationary solution is stable in one network topology and unstable in another topology. We thus repeat the sampling procedure above, but draw two topologies for every $\delta, \ensuremath{k}$ value pair. We discard all samples except those where the stationary state is stable in one topology and unstable in the other. Since we already know that the stability transition at the bottom bifurcation line is narrow for large $\ensuremath{k}$ we sample this region selectively by drawing $\delta$ from a bounded uniform distribution centered on $\delta_{d=3}(\ensuremath{k})$, the bifurcation point for DHONs with degree $d=3$, which is given by Eq.~\eqref{eq:delta_approx}. This results in a uniform sampling of the parameter plane shown in Fig.~\ref{fig:AvgDeg}. We observe that for large $\ensuremath{k}$ the transitions occur only at specific values of $\delta$. These values very closely approximate the analytical predictions of the bifurcation points in DHONs, where we replaced the homogeneous degree by the mean degree $d=2K/N$ of the respective DHEN. For large coupling strength the bifurcation line that governs the onset of amplitude death at small coupling delays thus seems to depend only on the mean degree of the network, whereas all other topological properties, at least in the Erd\H{o}s-Reny\'{i} ensemble, can be neglected. \section{Conclusions} In the present paper we investigated amplitude death in general networks for delay-coupled delay oscillators. Building on a previous result we were able to study the regions for amplitude death analytically for degree homogeneous networks. In particular we considered the limit of large coupling strength for which we showed that all relevant bifurcation lines can be traced back to certain eigenvalues of the systems adjacency matrix. We showed analytically that in degree homogeneous networks regions of amplitude death exist at large coupling strength $k$. Specifically, we investigated a region for which the onset of amplitude death occurs already at very small values of the coupling delay that scales as $1/k$. The width of this region scales as $1/\sqrt{k}$ and thus is significant, at least in a logarithmic sense. It is remarkable that this region of amplitude death comes arbitrarily close to, but never reaches the case of undelayed coupling. The results obtained here for delay oscillators contrast with the well-studied case of Stuart-Landau oscillators \cite{Reddy1998}, which do not show amplitude death for arbitrary large coupling strength and arbitrary small coupling delays. We can speculate that the strong self-feedback in delay oscillators might be in destructive resonance with the terms arising from the coupling to neighboring nodes. Thus, in contrast to Stuart-Landau oscillators a direct force working against the oscillatory dynamics might not be necessary. Numerical investigations indicate that the analytical results for degree homogeneous networks can be extended to the case of degree heterogeneous networks. For sufficiently large coupling strength we found that numerical results for degree heterogeneous networks were in perfect agreement with analytical expectations for degree homogeneous networks. While these results were based on network topologies from an ensemble of Erd\H{o}s-Reny\'{i} random graphs one can assume that they should hold generally, at least for networks with exponentially decaying degree distribution. Regarding degree heterogeneous networks we noted that the onset of amplitude death for strong coupling and small coupling delays seems to depend on the networks mean degree, but not on other properties. This is remarkable as it shows that the node dynamics become sensitive to a global and hence delocalized quantity, which hints at a diverging correlation length. This can be made plausible by considering that strong coupling creates a stiff system that can rapidly communicate over the relatively small diameter of the networks. Furthermore, we have seen in the degree homogeneous networks that the dominating instability in the limit of strong coupling are in-phase oscillations, which are an inherently global phenomenon.
\section{Introduction} The coagulation process describes the kinetics of particle growth where particles can coagulate to form larger particles via binary interaction. On the other side, the fragmentation process describes how particles break into two or more fragments. Examples of these processes can be found e.g. in astrophysics, polymer science \cite{ziff}, and cloud physics \cite[Chapter 15]{Pruppacher}. The dynamic of the coagulation-fragmentation process is described by the integro-differential equation \begin{eqnarray} &&\frac{\partial u(x,t)}{\partial t}=\frac{1}{2}\int\limits_0^x K(x-y,y)u(x-y,t)u(y,t)\,dy-\int\limits_0^\infty K(x,y)u(x,t)u(y,t)\,dy \nonumber\\ &&\qquad\quad\qquad+\int\limits_x^\infty b(x,y)S(y)u(y,t)\,dy-S(x)u(x,t), \label{Fproblema} \end{eqnarray} with initial condition \begin{eqnarray} u(x,0)=u_0(x)\geq 0\quad \mbox{a.e.} \label{Fcondinicial} \end{eqnarray} where the non-negative variables $x$ and $t$ represent the size of the particles and time respectively. The values $u(x,t)$ denote the number density of particles with size $x$ at time $t$. The rate at which particles of size $x$ coalesce with particles of size $y$ is represented by the coagulation kernel $K(x,y)$. The rate at which particules of size $x$ are selected to break is determined by the selection function $S(x)$. The breakage function $b(x,y)$ gives the number of particules of size $x$ produced when a particule of size $y$ breaks up. The equation (\ref{Fproblema}) is named the continuous coagulation equation with multifragmentation where the multifragmentation kernel $\Gamma$ defines the selection function $S$ and the breakage function $b$ by \begin{eqnarray} S(x)=\int\limits_0^x\frac{y}{x}\Gamma(x,y)dy, \qquad\quad b(x,y)=\Gamma(y,x)/S(y), \label{FB} \end{eqnarray} or vice versa. The breakage function is assumed here to have the following properties \begin{eqnarray} \int\limits_0^yxb(x,y)dx=y\quad\mbox{for all}\quad y>0, \label{Fb} \end{eqnarray} which is the conservation of mass and \begin{eqnarray} \int\limits_0^yb(x,y)dx=N<\infty\quad\mbox{for all}\quad y>0,\; b(x,y)=0\quad\mbox{for}\quad x>y, \label{FbN} \end{eqnarray} where the parameter $N$ represents the number of particles produced in fragmentation events. In this paper $N$ is assumed to be finite and independent of $y$. Equation (\ref{Fb}) allows the system to conserve the total mass during the fragmentation events. It states that the total mass of the fragments is equal to the mass $y$ of the particle that breaks. The existence and uniqueness of solutions to the coagulation-fragmentation equation has already been the subject of several papers. The case of multifragmentation, that is, when the particules can break into two or more parts, has also been studied, see e.g.\ \cite{PhLaurencot}, \cite{McLaughlin}, \cite{Melzak}, and \cite{Walker}. For more recent result see e.g.\ \cite{BanasiakLamb}, \cite{LambBanasiak}, \cite{Ankikumar} and \cite{AnkikWarnecke}. Giri et al.\cite{Ankikumar} studied the coagulation kernels of the form $K(x,y)=\phi(x)\phi(y)$ for some sublinear function $\phi$ under the growth restriction $\phi(x)\leq(1+x)^\mu$ for $0\leq\mu<1$, and the selection function $S(x)$ is there also considered under the same growth assumption. In \cite{AnkikWarnecke}, Giri et al. proved the existence of solutions to the coagulation equation with multifragmentation for a more general fragmentation kernel, in order to cover the fragmentation kernel $\Gamma(y,x)=(\alpha + 2)x^\alpha y^{\gamma-(\alpha+1)}$ getting a result for $\alpha>-1$ and $\gamma\in]0,\alpha+2[$. The existence proofs in \cite{AnkikWarnecke} and \cite{Ankikumar} are based on the well known basic method by Stewart \cite{Stewart}, where the solution is obtained through the convergence of the solutions to a sequence of truncated problems. In \cite{AnkikWarnecke} the uniqueness of the solutions was not studied. In \cite{BanasiakLamb} Banasiak and Lamb proved the existence and uniqueness of solutions to the coagulation-fragmentation equation when $K(x,y)\in L_\infty(\mathbb{R}_+\times\mathbb{R}_+)$ while in \cite{LambBanasiak} the authors proved existence and uniqueness of classical solutions for the class of coagulation kernels $K(x,y)\leq k\big((1+a(x))^\alpha+(1+a(y))^\alpha\big)$ where $a$ is the fragmentation rate, $k>0$, and $0\leq\alpha<1$. To our knowledge there is just one result concerning the coagulation mutlifragmentation equation with singular coagulation kernels. Ca\~nizo Rinc\'on \cite{Canizo} proved the existence of $L^\infty\big([0,T[,M_1\big)$ solutions in the distribution sense for the coagulation kernels $a(y,y')$ such that \begin{eqnarray*} K_a\big(y^\alpha(y')^\beta+(y')^\alpha y^\beta\big)\leq a(y,y')\leq K'_a\big(y^\alpha(y')^\beta+(y')^\alpha y^\beta\big) \end{eqnarray*} with $\alpha<\beta<1$, $0<\alpha+\beta<1$, $\beta-\alpha<1$, constants $K_a,K'_a>0$, and where $M_1$ is the space of measures $\mu$ on $(0,\infty)$ with first bounded moment, see \cite[Section 3.2]{Canizo} or \cite[page 59]{Canizo}. His result is resticted to the kernels with order $\alpha$ and $\beta$ in $y$ and $y'$ respectively, and being $\alpha\neq\beta$. The singularity is restricted to the case $\sigma\in]0,1/2[$ translated into our terms. The result from \cite{Canizo} leaves out, for example, the cases of the Smoluchowski and the equi-partition of kinetic energy kernels. The uniqueness of the solutions was not studied in \cite{Canizo}. In the present article, our aim is to prove the existence and uniqueness of solutions to the coagulation equation with multifragmentation with singular coagulation kernels \begin{eqnarray} K(x,y)\leq k(1+x)^\lambda(1+y)^\lambda(xy)^{-\sigma} \quad\mbox{with}\quad\lambda-\sigma\in[0,1[, \sigma\geq 0, \label{Fsmolokernel} \end{eqnarray} giving in this way an existence and uniqueness result for the case of the important Smoluchowki coagulation kernel \begin{eqnarray*} K(x,y)=(x^{1/3}+y^{1/3})(x^{-1/3}+y^{-1/3}) \end{eqnarray*} for Brownian motion, see Smoluchowski \cite{Smoluchowski}. The equi-partition of kinetic energy (EKE) kernel \begin{eqnarray*} K(x,y)=(x^{1/3}+y^{1/3})^2\sqrt{\frac{1}{x}+\frac{1}{y}}, \end{eqnarray*} is also covered by our analysis. We are giving a more general result than Ca\~nizo Rinc\'on \cite{Canizo} since we do not restrict our kernel to an specific order. We allow $\alpha$ to be equal $\beta$, the singularity can be as big as it is wished, and we obtain \emph{regular} solutions in the space $C^1\left([0,T],L^1\big(]0,\infty[\big)\right)$ with $T>0$. Our existence result is based on the proof of Stewart \cite{Stewart}. We extend the methods we developed in \cite{Carlos} for singular kernels in the pure coagulation problem. For our existence and uniqueness result we consider the class of fragmentation kernels \begin{eqnarray} \Gamma(y,x)\leq y^\theta b(x,y) \label{Ffragkernel} \end{eqnarray} with \begin{eqnarray} \int\limits_0^yb(x,y)x^{-2\sigma}\leq Cy^{-2\sigma}\quad\mbox{for}\quad \theta\in[0,1[, \sigma\in[0,1/2],\;\mbox{and a constant}\; C \label{Ffragkernel1} \end{eqnarray} and such that, there exist $q>1$ and $\tau_1,\tau_2\in[-2\sigma-\theta,1-\theta]$ such that \begin{eqnarray} \int\limits_0^yb^q(x,y)\leq B_1y^{q\tau_1},\;\mbox{and}\;\int\limits_0^yx^{-q\sigma}b^q(x,y)\leq B_2y^{q\tau_2}\quad\mbox{for constant $B_1,B_2>0$.} \label{Ffragkernel2} \end{eqnarray} From (\ref{FB}) and (\ref{Ffragkernel}) we have that $S(y)\leq y^\theta$. The case $S(y)=y^\theta$ with $\theta>0$ was considered in \cite{McGuinness}, where McGuinness et al.\ studied the pure fragmentation equation with singular initial conditions. The selection function $S(y)=y^\theta$ has also been studied in \cite{LambBrideGuinness}, \cite{RMZiff}, and \cite{ZiffGrady}. In \cite{McGrady} it has been considered for $\theta=0$. The class of frangmentation kernels (\ref{Ffragkernel}) holding (\ref{Ffragkernel1}) and (\ref{Ffragkernel2}) includes the kernel $\Gamma(y,x)=(\alpha + 2)x^\alpha y^{\gamma-(\alpha+1)}$ for $\alpha>2\sigma+\epsilon-1$ and $\gamma\in]0,1[$ with $0<\epsilon<\theta$. This kernel was studied by Giri et al. \cite{AnkikWarnecke}, where they proved the existence of weak solutions to the coagulation equation with multifragmentation, but with nonsingular coagulation kernels. In order to study the existence of solutions of (\ref{Fproblema})-(\ref{Fcondinicial}), we define for some given $\sigma\geq 0$ the space $Y$ to be the following Banach space with norm $\|\cdot\|_Y$ \begin{eqnarray*} Y=\left\{u\in L^1(]0,\infty[): \|u\|_Y<\infty\right\}\quad\mbox{where}\quad\|u\|_Y=\int\limits_0^\infty(x+x^{-2\sigma})|u(x)|dx. \end{eqnarray*} That $Y$ is a Banach space is easily seen. We also write \begin{eqnarray*} \|u\|_x=\int\limits_0^\infty x|u(x)|dx\quad\mbox{and}\quad\|u\|_{x^{-2\sigma}}=\int\limits_0^\infty x^{-2\sigma}|u(x)|dx, \end{eqnarray*} and set \begin{eqnarray*} Y^+=\left\{u\in Y:u\geq 0 \quad a.e.\right\}. \end{eqnarray*} Now we define a weak solution to problem (\ref{Fproblema})-(\ref{Fcondinicial}) in the same way as Stewart \cite{Stewart}: \begin{dfn} Let $T\in]0,\infty]$. A solution $u(x,t)$ of (\ref{Fproblema})-(\ref{Fcondinicial}) is a function $u:[0,T[\longrightarrow Y^+$ such that for a.e. $x\in[0,\infty[$ and $t\in[0,T[$ the following properties hold \begin{description} \item[(i)]$u(x,t)\geq 0$ for all $t\in[0,\infty[$, \item[(ii)]$u(x,\cdot)$ is continuous on $[0,T[$, \item[(iii)]for all $t\in[0,T[$ the following integral is bounded \begin{eqnarray*} \int\limits_0^t\int\limits_0^\infty K(x,y)u(y,\tau)\,dy\,d\tau<\infty\quad\mbox{and}\quad\int\limits_0^t\int\limits_x^\infty b(x,y)S(y)u(y,\tau)\,dy\,d\tau<\infty \end{eqnarray*} \item[(iv)] for all $t\in[0,T[$, $u$ satisfies the following weak formulation of (\ref{Fproblema}) \begin{eqnarray} u(x,t)\!\!&\!\!=\!\!&\!\!u(x,0)+\int\limits_0^t\left[\frac{1}{2}\int\limits_0^xK(x-y,y)u(x-y,\tau)u(y,\tau)\,dy-\int\limits_0^\infty K(x,y)u(x,\tau)u(y,\tau)\,dy\right. \nonumber\\ \!\!&\!\!\!\!&\!\!\qquad\qquad\qquad\left.+\int\limits_x^\infty b(x,y)S(y)u(y,t)\,dy-S(x)u(x,t)\right]d\tau.\label{elgranasterisco} \end{eqnarray} \end{description} \label{Fdefinicion} \end{dfn} In the next sections we make use of the following hypotheses \begin{hyp} \begin{description} \item[] \item[(H1)] $K(x,y)$ is a continuous non-negative function on $]0,\infty[\times]0,\infty[$, \item[(H2)] $K(x,y)$ is a symmetric function, i.e. $K(x,y)=K(y,x)$ for all $x,y\in]0,\infty[$, \item[(H3)] $K(x,y)\leq \kappa(1+x)^\lambda(1+y)^\lambda(xy)^{-\sigma}$ for $\lambda-\sigma\in[0,1[,\sigma\geq 0$, and constant $\kappa$, \item[(H4)] $S(x):]0,\infty[\rightarrow [0,\infty[$ is a continuous non-negative function such that $0\leq S(x)\leq y^\theta$ for $\theta\in[0,1[$, \item[(H5)] $b(x,y)$ is such that $\int\limits_0^yb(x,y)x^{-2\sigma}dx\leq Cy^{-2\sigma}$, \item[(H6)] there exist $q>1$ and $\tau_1,\tau_2\in[-2\sigma-\theta,1-\theta]$ such that \begin{eqnarray*} \int\limits_0^yb^q(x,y)\leq B_1y^{q\tau_1},\;\mbox{and}\;\int\limits_0^yx^{-q\sigma}b^q(x,y)\leq B_2y^{q\tau_2}\quad\mbox{for constant $B_1,B_2>0$} \end{eqnarray*} \end{description} \label{Fhypo} \end{hyp} In the rest of the paper we consider $\kappa=1$ for the simplicity. We study the uniqueness of the solutions to (\ref{Fproblema})-(\ref{Fcondinicial}) under the following further hypotheses \begin{description} \item[(H3')] $K(x,y)\leq\kappa_1(x^{-\sigma}+x^{\lambda-\sigma})(y^{-\sigma}+y^{\lambda-\sigma})$ \textit{such that} $\sigma\geq 0$, $\lambda-\sigma\in[0,1/2]$, \textit{and} $\kappa_1>0$, \item[(H4')] $S(x):]0,\infty[\rightarrow [0,\infty[$ is a continuous non-negative function such that $S(x)=y^\theta$ \textit{for} $\theta\leq\lambda-\sigma$. \end{description} The restriction $\lambda-\sigma\in[0,1/2]$ in \textbf{(H3')} limits our uniqueness result to a subset of the kernels of the class defined in \textbf{(H3)}, namely to the ones for which $\lambda-\sigma\in[0,1/2]$ holds. The restriction $\theta\leq\lambda-\sigma$ in \textbf{(H4')} limits our uniqueness result to a more restricted class of fragmentation kernels. We introduce now some easily derived inequalities that will be used throughout the paper. The proof of these inequalities can be found in Giri \cite{AnkikThesis}. For any $x,y>0$ \begin{eqnarray} 2^{p-1}(x^p + y^p)\leq (x + y)^p \leq x^p + y^p\quad &\mbox{if} &\quad 0\leq p\leq 1, \label{Finequality1}\\ 2^{p-1}(x^p + y^p)\geq (x + y)^p\geq x^p + y^p\quad &\mbox{if} &\quad p\geq 1, \label{Finequality2}\\ 2^{p-1}(x^p + y^p)\geq (x + y)^p\quad &\mbox{if} &\quad p < 0. \label{Finequality3} \end{eqnarray} The paper is organized as follows. In Section $2$ we define the sequence of truncated problems and prove in Theorem \ref{Fexistencetruncated} the existence and uniqueness of solutions to them. We extract a weakly convergent subsequence in $L^1$ from a sequence of unique solutions for truncated equations to (\ref{Fproblema})-(\ref{Fcondinicial}). In Section $3$ we show that the solution of (\ref{Fproblema}) is actually the limit function obtained from the weakly convergent subsequence of solutions of the truncated problem. In Section $4$ we prove the uniqueness, based on the method of Stewart \cite{Stewartuniq}, of the solutions to (\ref{Fproblema})-(\ref{Fcondinicial}) for a modification of the classes (\ref{Fsmolokernel}) and (\ref{Ffragkernel}) of coagulation and fragmentation kernels respectively. We obtain uniqueness for some kernels which are not covered by the existence result. \section{The Truncated Problem} $\quad$ We prove the existence of a solution to the problem (\ref{Fproblema})-(\ref{Fcondinicial}) by taking the limit of the sequence of solutions of the equations given by replacing the kernel $K(x,y)$ and the selection function $S(x)$ by their respective 'cut-off' kernel $K_n(x,y)$ and $S_n(x)$ for any given $n\in\mathbb{N}$ \begin{eqnarray} \begin{array}{ll} K_n(x,y)=\left\{ \begin{array}{ll} K(x,y) & \mbox{if}\quad x+y\leq n\;\;\mbox{and}\;\; x,y\geq \sigma/n \\ 0 & \mbox{otherwise}. \end{array} \right. & S_n(x)=\left\{ \begin{array}{ll} S(x) & \mbox{if}\quad x\leq n \\ 0 & \mbox{otherwise}. \end{array} \right. \end{array} \label{FKnSn} \end{eqnarray} For the defined kernels the resulting equations are written as \begin{eqnarray} \dfrac{\partial u^n(x,t)}{\partial t}\!\!&\!\!=\!\!&\!\!\dfrac{1}{2}\int\limits_0^x{K_n(x-y,y)u^n(x-y,t)u^n(y,t)}\,dy - \int\limits_0^{n-x}{K_n(x,y)u^n(x,t)u^n(y,t)}\,dy \nonumber\\ \!\!&\!\!\!\!&\!\!+\int\limits_x^nS_n(y)b(x,y)u^n(y,t)\,dy - S_n(x)u^n(x,t), \label{Fcen} \end{eqnarray} with the truncated initial data \begin{eqnarray} u^n_0(x)=\left\{ \begin{array}{ll} u_0(x) & \mbox{if}\quad 0\leq x\leq n \\ 0 & \mbox{otherwise}, \end{array} \right. \label{Ficen1} \end{eqnarray} where $u^n$ denotes the solution of the problem (\ref{Fcen})-(\ref{Ficen1}) for $x\in[0,n]$. \begin{thm} Suppose that \textbf{(H1)-(H6)} hold and $u_0\in Y^+$. Then for each $n=2,3,4,\ldots$ the problem (\ref{Fcen})-(\ref{Ficen1}) has a unique solution $u^n$ with $u^n(x,t)\geq 0$ for a.e. $x\in[0,n]$ and $t\in [0,\infty[$. Moreover, for all $t\in [0,\infty[$ \begin{equation} \int_0^n{xu^n(x,t)}\,dx=\int_0^n{xu^n(x,0)}\,dx. \label{Fconservation} \end{equation} \label{Fexistencetruncated} \end{thm} The proof of Theorem \ref{Fexistencetruncated} follows proceeding as in \cite[Theorem 3.1]{Stewart}. \subsection{Properties of the solutions of the truncated problem} \begin{lem} Let $u^n$ a solution of the truncated problem (\ref{Fcen})-(\ref{Ficen1}). Then for $\alpha\geq 0$ and $n=1,2,\ldots$ we obtain the inequality \begin{eqnarray*} \frac{d}{dt}\int\limits_0^nu^n(x,t)x^{-\alpha}dx\leq\int\limits_0^n\int\limits_0^yS_n(y)b(x,y)u^n(y,t)x^{-\alpha}dx\,dy. \end{eqnarray*} \label{Fdesigualdad} \end{lem} \textbf{Proof.} Multiplying equation (\ref{Fcen}) by $x^{-\alpha}$ and integrating w.r.t $x$ from $0$ to $n$, changing the order of integration, then a change of variable $x-y=z$, and again re-changing the order of integration while replacing $z$ by $x$ gives \begin{eqnarray} \frac{d}{dt}\int\limits_0^nu^n(x,t)x^{-\alpha}dx\!\!&\!\!=\!\!&\!\!\frac{1}{2}\int\limits_0^n\int\limits_0^{n-x}K_n(x,y)u^n(x,t)u^n(y,t)\left[(x+y)^{-\alpha}-x^{-\alpha}-y^{-\alpha}\right]dy\,dx \nonumber\\ \!\!&\!\!\!\!&\!\! +\int\limits_0^n\int\limits_x^nS_n(y)b(x,y)u^n(y,t)x^{-\alpha}dy\,dx-\int\limits_0^nS_n(x)u^n(x,t)x^{-\alpha}dx.\label{22nuevo} \end{eqnarray} Using the fact that $x^{-\sigma}$ is a sublinear function and therefore $(x+y)^{-\alpha}-x^{-\alpha}-y^{-\alpha}\leq 0$ we get \begin{eqnarray*} \frac{d}{dt}\int\limits_0^nu^n(x,t)x^{-\alpha}dx\leq\int\limits_0^n\int\limits_x^nS_n(y)b(x,y)u^n(y,t)x^{-\alpha}dy\,dx=\int\limits_0^n\int\limits_0^yS_n(y)b(x,y)u^n(y,t)x^{-\alpha}dx\,dy, \end{eqnarray*} which complete the proof of the theorem.\hfill{$\Box$} In the rest of the paper we consider for each $u^n$ their zero extension on $\mathbb R$, i.e. \begin{eqnarray*} \hat{u}^n(x,t)=\left\{ \begin{array}{ll} u^n(x,t) & 0\leq x\leq n, \;\; t\in[0,T],\\ 0 & x<0\;\;\mbox{or}\;\;x>n. \end{array} \right. \end{eqnarray*} For clarity we drop the notation $\hat{\cdot}$ for the remainder of the paper. \begin{lem} Assume that \textbf{(H1)-(H6)} hold. We take $u^n$ to be the non-negative zero extension of the solution to the truncated problem found in Theorem \ref{Fexistencetruncated}. Fix $T>0$ and let us define $$L(T)=\left(e^{NT}(N+1)+e^{CT}(C+1)+1\right)\|u_0\|_Y.$$ Then the following are true: \begin{description} \item[(i)]We have the bound \begin{eqnarray*}\int\limits_0^\infty{(1+x+x^{-2\sigma})u^n(x,t)}\,dx \leq L(T)\quad\mbox{for all}\quad t\in[0,T]. \end{eqnarray*} \item[(ii)]Given $\epsilon>0$ there exists an $R>1$ such that for all $t\in[0,T]$ \begin{eqnarray*} \sup_n\left\{\int\limits_R^\infty(1+x^{-\sigma})u^n(x,t)\,dx\right\}\leq\epsilon. \end{eqnarray*} \item[(iii)]Given $\epsilon>0$ there exists a $\delta>0$ such that for all $n=2,3,\ldots$ and $t\in\left[0,T\right]$ \begin{eqnarray} \int\limits_A (1+x^{-\sigma})u^n(x,t)\,dx<\epsilon \quad\quad\mbox{whenever}\quad\quad \mu(A)<\delta, \label{prop3} \end{eqnarray} where $\mu(\cdot)$ denotes the Lebesgue measure. \end{description} \label{Flemaproperty} \end{lem} \textbf{Proof.}$\;$\textbf{Property (i)} Computing the term with the weight $x^{-2\sigma}$ using Lemma \ref{Fdesigualdad} for $\alpha=2\sigma$ and using \textbf{(H5)} we get \begin{eqnarray*} \frac{d}{dt}\int\limits_0^nu^n(x,t)x^{-2\sigma}dx\!\!&\!\!\leq\!\!&\!\!\int\limits_0^n\int\limits_0^yS(y)b(x,y)u^n(y,t)x^{-2\sigma}dx\,dy \\ \!\!&\!\!\leq\!\!&\!\!C\int\limits_0^ny^{\theta-2\sigma}u^n(y,t)\,dy \\ \!\!&\!\!\leq\!\!&\!\!C\int\limits_0^1y^{-2\sigma}u^n(y,t)\,dy +C\int\limits_1^nyu^n(y,t)\,dy\\ \!\!&\!\!\leq\!\!&\!\!C\int\limits_0^nx^{-2\sigma}u^n(x,t)\,dx +C\|u_0\|_Y. \end{eqnarray*} From this inequality we find as above that \begin{eqnarray} \int\limits_0^nu^n(x,t)x^{-2\sigma}dx\leq e^{Ct}(C+1)\|u_0\|_Y, \qquad t\in[0,T]. \label{F2} \end{eqnarray} Now, since $1\leq x+x^{-2\sigma}$, by the mass conservation property (\ref{Fconservation}) and by (\ref{F2}) we obtain \begin{eqnarray*} \int\limits_0^\infty(1+x+x^{-2\sigma})u^n(x,t)\;dx\!\!&\!\!\leq\!\!&\!\!2\int\limits_0^n(x+x^{-2\sigma})u^n(x,t)\;dx\\ \!\!&\!\!\leq\!\!&\!\!2\left(\|u_0\|_Y+e^{Ct}(C+1)\|u_0\|_Y\right) \\ \!\!&\!\!\leq\!\!&\!\!2\left(e^{CT}(C+1)+1\right)\|u_0\|_Y=: L(T). \end{eqnarray*} \textbf{Property (ii)} Choose $\epsilon >0$ and let $R>1$ be such that $R>\frac{2\|u_0\|_Y}{\epsilon}$. Then using (\ref{Fconservation}) we get \begin{eqnarray*} \int\limits_R^\infty(1+x^{-\sigma})u^n(x,t)\,dx = 2\int\limits_R^\infty\frac{x}{x}u^n(x,t)\,dx \leq \frac{2}{R}\int\limits_0^n xu^n(x,t)\,dx\leq\frac{2}{R}\|u^n_0\|_Y\leq\frac{2}{R}\|u_0\|_Y<\epsilon. \end{eqnarray*} \textbf{Property (iii)} Let $\chi_A$ denote the characteristic function of a set $A$ and set \begin{eqnarray} \kappa(r):=\frac{1}{2}(1+r^\sigma)(1+r)^{2\lambda}. \label{Fkappa} \end{eqnarray} Let us define for all $n=1,2,3,\ldots$ and $t\in[0,T]$ using property \textbf{(i)} \begin{eqnarray} f^n(\delta,r,t)=\sup\left\{\int\limits_0^r\chi_{A}(x)(1+x^{-\sigma})u^n(x,t)\,dx:\; A\subset]0,r[\;\mbox{and}\;\mu(A)<\delta\right\}\leq L(T). \label{Fcond10} \end{eqnarray} In the rest of the paper we drop the argument $r$ from $f^n$ for simplicity. We take $t=0$ in the definition of $f^n$ and observe that $u^n(x,0)\leq u_0(x)$ pointwise almost everywhere. Then by the absolute continuity of the Lebesgue integral, we have that \begin{eqnarray} f^n(\delta,0)=\sup\left\{\int\limits_0^r\chi_{A}(x)(1+x^{-\sigma})u_0^n(x)\,dx:\; A\subset]0,r[\;\mbox{and}\;\mu(A)<\delta\right\}\rightarrow 0\;\mbox{as $\delta\rightarrow 0$} \label{Fcond1} \end{eqnarray} Now we multiply (\ref{Fcen}) by $(1+x^{-\sigma})\chi_{A}(x)$. This we integrate from $0$ to $t$ w.r.t. $s$ and over $[0,r[$ w.r.t. $x$. Using the non-negativity of each $u^n$ we obtain \begin{eqnarray} \!\!&\!\!\!\!&\!\!\int\limits_0^r\chi_{A}(x)(1+x^{-\sigma})u^n(x,t)\,dx \nonumber\\ \!\!&\!\!\!\!&\!\!\qquad\leq\frac{1}{2}\int\limits_0^t\int\limits_0^r\int\limits_0^x\chi_{A}(x)(1+x^{-\sigma})K_n(x-y,y)u^n(x-y,s)u^n(y,s)\,dy\,dx\,ds\label{diferenciapropiedad}\\ \!\!&\!\!\!\!&\!\!\qquad\quad+ \int\limits_0^t\int\limits_0^r\chi_{A}(x)\int\limits_x^nS_n(y)b(x,y)(1+x^{-\sigma})u^n(y,s)\,dy\,dx\,ds + \int\limits_0^r\chi_{A}(x)(1+x^{-\sigma})u^n_0(x)\,dx.\nonumber \end{eqnarray} Let us denote $I_{21}$ and $I_{22}$ the first and the second integral terms on the right hand side of (\ref{diferenciapropiedad}) respectively. By changing variables as we did in (\ref{22nuevo}) in $I_{21}$ we get \begin{eqnarray*} I_{21}(t)=\frac{1}{2}\int\limits_0^t\int\limits_0^r\int\limits_0^{r-y}\chi_{A}(x+y)[1+(x+y)^{-\sigma}]K_n(x,y)u^n(x,s)u^n(y,s)\,dx\,dy\,ds. \nonumber\\ \end{eqnarray*} By using \textbf{(H3)} for $K(x,y)$, then taking $1+(x+y)^{-\sigma}\leq 1+y^{-\sigma}$ and $x^{-\sigma}\leq 1+x^{-\sigma}$ we have \begin{eqnarray*} I_{21}(t)\!\!&\!\!\leq\!\!&\!\!\frac{1}{2}\int\limits_0^t\int\limits_0^r\int\limits_0^{r-y}\chi_{A}(x+y)\left[1+(x+y)^{-\sigma}\right](1+x )^\lambda(1+y)^\lambda(xy)^{-\sigma}u^n(x,s)u^n(y,s)\,dx\,dy\,ds \\ \!\!&\!\!\leq\!\!&\!\!\frac{1}{2}\int\limits_0^t\int\limits_0^r\int\limits_0^{r-y}\chi_{A}(x+y)(1+y^{-\sigma})(1+x )^\lambda(1+y)^\lambda(1+x^{-\sigma})y^{-\sigma} u^n(x,s)u^n(y,s)\,dx\,dy\,ds \\ \!\!&\!\!=\!\!&\!\!\frac{1}{2}\int\limits_0^t\int\limits_0^{r}\int\limits_0^{r-y}\chi_{A}(x+y)(1+y^{\sigma})(1+x )^\lambda(1+y)^\lambda(1+x^{-\sigma})y^{-2\sigma} u^n(x,s)u^n(y,s)\,dx\,dy\,ds. \end{eqnarray*} By using the definition (\ref{Fkappa}) of $\kappa(r)$ we obtain the following estimates for $I_{21}$ \begin{eqnarray*} I_{21}(t)\!\!&\!\!\leq\!\!&\!\! \kappa(r)\int\limits_0^t\int\limits_0^ru^n(y,s)y^{-2\sigma}\int\limits_0^\infty\chi_{A-y\cap[0,r-y]}(x)(1+x^{-\sigma})u^n(x,s)\,dx\,dy\,ds \nonumber\\ \end{eqnarray*} where $A-y:=\left\{\omega>0:\omega=x-y\,\mbox{for some}\,x\in A\right\}$. Since $A-y\cap[0,r-y]\subset [0,r]$ and $\mu\left(A-y\cap[0,r-y]\right)\leq\mu(A-y)\leq\mu(A)<\delta$, by using the definition of $f^n$ and property \textbf{(i)} we have \begin{eqnarray} I_{21}(t)\leq \kappa(r)L(T)\int\limits_0^tf^n(\delta,s)\,ds. \label{I21*} \end{eqnarray} Working now with the integral term $I_{22}$ we have using \textbf{(H4)} that \begin{eqnarray*} I_{22}(t)\!\!&\!\!\leq\!\!&\!\!\int\limits_0^t\int\limits_0^r\chi_{A}(x)\int\limits_x^\infty S_n(y)b(x,y)(1+x^{-\sigma})u^n(y,s)\,dy\,dx\,ds \nonumber\\ &\leq&\int\limits_0^t\int\limits_0^\infty\int\limits_0^y\chi_A(x)S_n(y)b(x,y)(1+x^{-\sigma})u^n(y,s)\,dx\,dy\,ds\nonumber\\ &\leq&\int\limits_0^t\int\limits_0^\infty y^{\theta}u^n(y,s)\left[\int\limits_0^y\chi_A(x)b(x,y)\,dx+\int\limits_0^y\chi_A(x)b(x,y)x^{-\sigma}dx\right]dy\,ds.\nonumber \end{eqnarray*} Then by hypotheses \textbf{(H6)} we find \begin{eqnarray*} I_{22}(t)\!\!&\!\!\leq\!\!&\!\!\mu(A)^{\frac{p-1}{p}}\int\limits_0^t\int\limits_0^\infty y^{\theta}u^n(y,s)\left[\left(\int\limits_0^yb^p(x,y)dx\right)^{1/p}+\left(\int\limits_0^yb^p(x,y)x^{-p\sigma}dx\right)^{1/p}\right]dy\,ds\\ &\leq&\mu(A)^{\frac{p-1}{p}}\int\limits_0^t\int\limits_0^\infty y^{\theta}u^n(y,s)\left(B_1y^{\tau_1}+B_2y^{\tau_2}\right)dy\,ds\leq (B_1+B_2)\mu(A)^{\frac{p-1}{p}}L(T)T. \end{eqnarray*} Using the estimates of $I_{21}(t)$ and $I_{22}(t)$ in (\ref{diferenciapropiedad}) we have by taking the supremum over all A such that $A\subset]0, r[$ with $\mu(A)\leq\delta$ \begin{eqnarray*} f^n(\delta,t)\leq\kappa(r)L(T)\int\limits_0^tf^n(\delta,s)ds+(B_1+B_2)L(T)T\delta^{\frac{p-1}{p}}+f^n(\delta,0),\quad t\in[0,T]. \end{eqnarray*} By using Gronwall's inequality, see e.g.\ Walter \cite[page 361]{Walter}, we get \begin{eqnarray} f^n(\delta,t)\leq\left[(B_1+B_2)L(T)T\delta^{\frac{p-1}{p}}+f^n(\delta,0)\right]\exp\big(\kappa(r)L(T)T\big),\quad t\in[0,T]. \label{Ffn} \end{eqnarray} Since $f^n(\delta,0)\rightarrow 0$ as $\delta\rightarrow 0$ (\ref{Ffn}) implies that \begin{eqnarray} \lim_{\delta\rightarrow 0}\sup_{n\geq 1, t\in[0,T]}\left\{f^n(\delta,t)\right\}=0. \label{propiedad} \end{eqnarray} Lemma \ref{Flemaproperty}\textbf{(iii)} is then a consequence of (\ref{propiedad}) and Lemma \ref{Flemaproperty}\textbf{(i)}. This completes the proof of Lemma \ref{Flemaproperty}. \hfill{$\Box$} Let us define $v^n(x,t)=x^{-\sigma}u^n(x,t)$. Due to the Lemma \ref{Flemaproperty} above and the Dunford-Pettis Theorem \cite[page 274]{Edwards}, we can conclude that for each $t\in[0,T]$ the sequences $\big(u^n(t)\big)_{n\in\mathbb N}$ and $\big(v^n(t)\big)_{n\in\mathbb N}$ are weakly relatively compact in $L^1\big(]0,\infty[\big)$. \subsection{Equicontinuity in time} \begin{lem} Assume that Hypotheses $1.1$ hold. Take $\big(u^n\big)$ now to be the sequence of extended solutions to the truncated problems (\ref{Fcen})-(\ref{Ficen1}) found in Theorem \ref{Fexistencetruncated} and $v^n(x,t)=x^{-\sigma}u^n(x,t)$. Then there exists a subsequences $\big(u^{n_k}(t)\big)$ and $\big(v^{n_l}(t)\big)$ of $\big(u^{n}(t)\big)_{n\in\mathbb N}$ and $\big(v^{n}(t)\big)_{n\in\mathbb N}$ respectively such that \begin{eqnarray*} && u^{n_k}(t)\rightharpoonup u(t)\quad\mbox{in}\quad L^1\big(]0,\infty[\big)\quad\mbox{as}\quad n_k\rightarrow\infty \\ && v^{n_l}(t)\rightharpoonup v(t)\quad\mbox{in}\quad L^1\big(]0,\infty[\big)\quad\mbox{as}\quad n_l\rightarrow\infty \end{eqnarray*} uniformly for $t\in [0,T]$. This convergence is uniform for all $t\in [0; T]$ giving $u,v\in C\left([0,T];\Omega\right)=\left\{\eta:[0,T]\rightarrow \Omega,\right.$ $\left.\eta \;\mbox{continuous and}\; \eta(t)\; \mbox{bounded for all}\; t\in[0,T] \right\}$, where $\Omega$ is $L^1\big(]0,\infty[\big)$ equipped with the weak topologyy.\label{Funlema} \end{lem} \textbf{Proof}: Choose $\epsilon>0$, and $\phi\in L^\infty\big(]0,\infty[\big)$. Let $s,t\in[0,T]$ and assume that $t\geq s$. Choose $a>1$ such that \begin{eqnarray} \frac{2L(T)}{a}\|\phi\|_{L^\infty(]0,\infty[)}\leq\epsilon/2. \label{F20} \end{eqnarray} Let us define the function $\omega(x,t):=u^n(x,t)x^{-\beta}$ for $\beta=0$ or $\beta=\sigma$. Note that for $\beta=0$ and $\beta=\sigma$ it becames $u^n(x,t)$ and $v^n(x,t)$ respectively. Using Lemma \ref{Flemaproperty}, for each $n$, we get using $a>1$ chosen to satisfy (\ref{F20}) \begin{eqnarray} \int\limits_a^\infty\left|\omega^n(x,t)-\omega^n(x,s)\right|dx\!\!&\!\!=\!\!&\!\!\int\limits_a^\infty\left|x^{-\beta}u^n(x,t)-x^{-\beta}u^n(x,s)\right|dx \nonumber\\ \!\!&\!\!\leq\!\!&\!\!\frac{1}{a}\int\limits_a^\infty x^{1-\beta}\left|u^n(x,t)+u^n(x,s)\right|dx\nonumber\\ \!\!&\!\!\leq\!\!&\!\!\frac{1}{a}\int\limits_a^\infty x\left|u^n(x,t)+u^n(x,s)\right|dx \leq 2L(T)/a. \label{Fotratu} \end{eqnarray} By using (\ref{Fcen}), (\ref{F20}), (\ref{Fotratu}), for $t\geq s$ we obtain \begin{eqnarray} \!\!&\!\!\!\!&\!\!\left|\int\limits_0^\infty\phi(x)\left[\omega^n(x,t)-\omega^n(x,s)\right]dx\right| \nonumber\\ \!\!&\!\!\!\!&\!\!\leq\int\limits_0^a|\phi(x)|\left|\omega^n(x,t)-\omega^n(x,s)\right|dx + \epsilon/2 \nonumber\\ \!\!&\!\!\!\!&\!\!\leq\|\phi\|_{L^\infty(]0,\infty[)}\int\limits_s^t\left[\frac{1}{2}\int\limits_0^a\int\limits_0^xK_n(x-y,y)u^n(x-y,\tau)u^n(y,\tau)x^{-\beta}dy\,dx\right. \\ \!\!&\!\!\!\!&\!\!\qquad\qquad\qquad\qquad\quad+\int\limits_0^a\int\limits_0^{n-x}K_n(x,y)u^n(x,\tau)u^n(y,\tau)x^{-\beta}dy\,dx \nonumber\\ \!\!&\!\!\!\!&\!\!\qquad\qquad\qquad\qquad\quad\left.+\int\limits_0^a\int\limits_x^nb(x,y)S_n(y)u^n(y,\tau)x^{-\beta}dy\,dx+\int\limits_0^aS_n(x)u^n(x,\tau)x^{-\beta}dx\right]d\tau + \epsilon/2 \nonumber\\ \!\!&\!\!\!\!&\!\!=\|\phi\|_{L^\infty(]0,\infty[)}\int\limits_s^t\left(I_{41}(\tau)+I_{42}(\tau)+I_{43}(\tau)+I_{44}(\tau)\right)d\tau +\epsilon/2. \nonumber \label{f28} \end{eqnarray} A change of variables in the first integral gives \begin{eqnarray*} I_{41}(\tau)=\frac{1}{2}\int\limits_0^a\int\limits_0^{a-x}K_n(x,y)u^n(x,\tau)u^n(y,\tau)(x+y)^{-\beta}dy\,dx. \end{eqnarray*} Taking $y=0$ in the term $(x+y)^{-\beta}$ we find that \begin{eqnarray} I_{41}(\tau)\leq \frac{1}{2}\int\limits_0^a\int\limits_0^{a-x}K_n(x,y)u^n(x,\tau)u^n(y,\tau)x^{-\beta}dy\,dx. \label{Farriba5} \end{eqnarray} By using the definition of $K_n(x,y)$, Lemma \ref{Flemaproperty}\textbf{(i)}, and the fact that $\beta$ just take the values $0$ and $\sigma$ we have \begin{eqnarray} I_{41}(\tau)\!\!&\!\!\leq\!\!&\!\!\frac{1}{2}\int\limits_0^a\int\limits_0^{a-x}(1+x)^\lambda(1+y)^\lambda x^{-(\sigma+\beta)}y^{-\sigma}u^n(x,\tau)u^n(x,\tau)\,dy\,dx \nonumber\\ \!\!&\!\!\leq\!\!&\!\! \frac{1}{2}(1+a)^{2\lambda}\int\limits_0^{a}\int\limits_0^{a-x}x^{-(\sigma+\beta)}y^{-\sigma}u^n(x,\tau)u^n(x,\tau)\,dy\,dx \leq\frac{1}{2}(1+a)^{2\lambda}L(T)^2. \label{FI41} \end{eqnarray} In order to estimate the second term, we define \begin{eqnarray} C_1=\left\{\begin{array}{lcl} 1 & \mbox{if} & 0\leq\lambda\leq 1 \\ 2^{\lambda-1} & \mbox{if} & \lambda\geq 1. \end{array} \right. \label{FC1} \end{eqnarray} Then, by using inequalities (\ref{Finequality1}) and (\ref{Finequality2}) for $p=\lambda$ and Lemma \ref{Flemaproperty}\textbf{(i)}, working as in (\ref{FI41}), we find that \begin{eqnarray} I_{42}(\tau)\!\!&\!\!\leq\!\!&\!\!C_1^2\int\limits_0^a\int\limits_0^{n-x}(1+x^\lambda)(1+y^\lambda)x^{-(\sigma+\beta)}y^{-\sigma}u^n(x,\tau)u^n(y,\tau)\,dy\,dx \nonumber\\ \!\!&\!\!\leq\!\!&\!\!C_1^2(1+a^\lambda)\int\limits_0^a\int\limits_0^{n-x}(y^{-\sigma}+y^{\lambda-\sigma})x^{-(\sigma+\beta)}u^n(x,\tau)u^n(y,\tau)\,dy\,dx \leq 2C_1^2(1+a^\lambda)L(T)^2. \label{FI42} \end{eqnarray} Now changing the order of integration in $I_{43}(\tau)$ we have \begin{eqnarray*} I_{43}(\tau)\!\!&\!\!\leq\!\!&\!\!\int\limits_0^a\int\limits_x^nb(x,y)S(y)u^n(y,\tau)x^{-\beta}dy\,dx \nonumber\\ \!\!&\!\!=\!\!&\!\!\int\limits_0^a\int\limits_0^yb(x,y)y^\theta u^n(y,\tau)x^{-\beta}dx\,dy+\int\limits_a^n\int\limits_0^ab(x,y)y^\theta u^n(y,\tau)x^{-\beta}dx\,dy. \label{F23*} \end{eqnarray*} We can see that, by using \textbf{(H5)} and (\ref{FbN}) for $\beta=\sigma$ \begin{eqnarray} \int\limits_0^yb(x,y)x^{-\beta}dx=\int\limits_0^yb(x,y)x^{-\sigma}dx\!\!&\!\!\leq\!\!&\!\!\int\limits_0^1b(x,y)x^{-2\sigma}dx + \int\limits_1^yb(x,y)dx \nonumber\\ \!\!&\!\!\leq\!\!&\!\!\int\limits_0^yb(x,y)x^{-2\sigma}dx + \int\limits_0^yb(x,y)dx \leq Cy^{-2\sigma}+N \label{Fprop3} \end{eqnarray} and for $\beta=0$ \begin{eqnarray} \int\limits_0^yb(x,y)x^{-\beta}dx=\int\limits_0^yb(x,y)dx=N. \label{Fprop33} \end{eqnarray} Then, from (\ref{Fprop3}) and (\ref{Fprop33}) we can conclude that \begin{eqnarray} \int\limits_0^yb(x,y)x^{-\beta}dx \leq Cy^{-2\sigma}+N \label{Fprop333} \end{eqnarray} By using (\ref{Fprop333}) and \textbf{(H4)}, $I_{43}(\tau)$ can be estimated by \begin{eqnarray} I_{43}(\tau)\!\!&\!\!\leq\!\!&\!\!\int\limits_0^a(Cy^{-2\sigma}+N)y^\theta u^n(y,\tau)\,dy+\int\limits_a^n(Cy^{-2\sigma}+N)y^\theta u^n(y,\tau)\,dy \nonumber\\ \!\!&\!\!=\!\!&\!\!\int\limits_0^n(Cy^{-2\sigma}+N)y^\theta u^n(y,\tau)\,dy\nonumber\\ \!\!&\!\!=\!\!&\!\!C\int\limits_0^ny^{\theta-2\sigma}u^n(y,\tau)\,dy+N\int\limits_0^ny^{\theta}u^n(y,\tau)\,dy\leq(C+N)L(T). \label{FI43} \end{eqnarray} Using \textbf{(H4)} and Lemma \ref{Flemaproperty}\textbf{(i)} we obtain \begin{eqnarray} I_{44}(\tau)\leq\int\limits_0^ax^{\theta-\beta} u^n(x,\tau)dx\leq L(T) \label{FI44} \end{eqnarray} which together with (\ref{FI41}), (\ref{FI42}) and (\ref{FI43}) brings (\ref{f28}) to \begin{eqnarray} &&\left|\int\limits_0^\infty\phi(x)\left[\omega^n(x,t)-\omega^n(x,s)\right]dx\right|\nonumber\\ &&\leq \left[\left(\frac{1}{2}(1+a)^{2\lambda}+2C_1^2\left(1+a^\lambda\right)\right)L(T)^2+(C+N+1)L(T)\right](t-s)\|\phi\|_{L^\infty(]0,\infty[)}+\epsilon/2<\epsilon,\label{34} \end{eqnarray} whenever $(t-s)<\delta$ for some $\delta>0$ sufficiently small. The argument given above similarly holds for $s<t$. Hence (\ref{34}) holds for all $n$ and $|t-s|<\delta$. Then the sequence $\big(\omega^n(t)\big)_{n\in\mathbb N}$ is time equicontinuous in $L^1\big(]0,\infty[\big)$.\ Thus, $\big(\omega^n(t)\big)$ lies in a relatively compact subset of a gauge space $\Omega$. The gauge space $\Omega$ is $L^1\big(]0,\infty[\big)$ equipped with the weak topology. For details about the gauge space, see Ash ~\cite[page 226]{Ash}. Then, we may apply a version of the Arzela-Ascoli Theorem, see Ash ~\cite[page 228]{Ash}, to conclude that there exists a subsequence $\big(\omega^{n_k}\big)_{k\in\mathbb N}$ such that \begin{eqnarray*} \omega^{n_k}(t)\rightarrow \omega(t)\quad\mbox{in}\quad \Omega\quad\mbox{as}\quad n_k\rightarrow\infty, \end{eqnarray*} uniformly for $t\in [0,T]$ for some $\omega\in C\left([0,T];\Omega\right)$. Then taking $\beta=0$ and $\beta=\sigma$ we can conclude that there exist subsequence $\big(u^{n_k}\big)_{k\in\mathbb N}$ and $\big(v^{n_k}\big)_{k\in\mathbb N}$ such that \begin{eqnarray*} u^{n_k}(t)\rightarrow u(t)\quad\mbox{in}\quad \Omega\quad\mbox{as}\quad n_k\rightarrow\infty\\ v^{n_k}(t)\rightarrow v(t)\quad\mbox{in}\quad \Omega\quad\mbox{as}\quad n_k\rightarrow\infty, \end{eqnarray*} uniformly for $t\in [0,T]$ for some $u,v\in C\left([0,T];\Omega\right)$. \hfill{$\Box$} \section{Existence Theorem} \subsection{Convergence of the integrals} In order to show that the limit function which we obtained above is indeed a solution to (\ref{Fproblema})-(\ref{Fcondinicial}), we define the operators $M^n_i$, $M_i$, $i=1,2,3,4$ \begin{eqnarray*} \begin{array}{ll} M_1^n(u^n)(x)=\frac{1}{2}\int\limits_{0}^xK_n(x-y,y)u^n(x-y)u^n(y)\,dy & M_1(u)(x)=\frac{1}{2}\int\limits_0^xK(x-y,y)u(x-y)u(y)\,dy \\ M_2^n(u^n)(x)=\int\limits_{0}^{n-x}K_n(x,y)u^n(x)u^n(y)\,dy & M_2(u)(x)=\int\limits_0^\infty K(x,y)u(x)u(y)\,dy \\ M_3^n(u^n)(x)=\int\limits_x^nb(x,y)S_n(y)u^n(y)\,dy & M_3(u)(x)=\int\limits_x^\infty b(x,y)S(y)u(y)\,dy \\ M_4^n(u^n)(x)=S_n(x)u^n(x) & M_4(u)(x)=S(x)u(x), \end{array} \end{eqnarray*} where $u\in L^1\big(]0,\infty[\big)$, $x\in]0,\infty[$ and $n=1, 2, \ldots$. Set $M^n=M^n_1-M^n_2+M^n_3-M^n_4$ and $M=M_1-M_2+M_3-M_4$. \begin{lem} Suppose that $\big(u^n\big)_{n\in\mathbb N}\subset Y^+$, $u\in Y^+$ where $\|u^n\|_Y\leq L$, $\|u\|_Y\leq Q$, $u^n\rightharpoonup u$ and $v^n\rightharpoonup v$ in $L^1(]0,\infty[)$ as $n\rightarrow\infty$, for $v=x^{-\sigma}u$ and $v^n=x^{-\sigma}u^n$. Then for each $a>0$ \begin{eqnarray*} M^n(u^n)\rightharpoonup M(u)\quad\mbox{in}\quad L^1\big(]0,a[\big)\quad\mbox{as}\quad n\rightarrow\infty. \end{eqnarray*} \label{Flemaconvergence} \end{lem} \textbf{Proof}: Choose $a>0$ and let $\phi\in L^\infty\big(]0,\infty[\big)$. We show that $M^n_i(u^n)\rightharpoonup M_i(u)$ in $L^1\big(]0,a[\big)$ as $n\rightarrow\infty$ for $i=1,2,3,4$. \\[14pt] The proof of case $i=1$ is analogous to the proof of the $W_1$ case in Stewart \cite[Lemma 4.1]{Stewart} by taking \begin{eqnarray*} h(v)(x)=\frac{1}{2}\int\limits_0^{a-x}\phi(x+y)K(x,y)(xy)^\sigma v(y)\,dy\quad\mbox{where}\quad v=x^{-\sigma}u. \end{eqnarray*} For every $\epsilon>0$ and $C_1$ defined by (\ref{FC1}) we can choose $\eta$ large enough, due to the negative exponents, such that for $L,Q$ from our assumptions \begin{eqnarray} C_1\|\phi\|_{L^\infty([0,a])}\left[(2\eta^{-(1+\sigma)}+\eta^{\lambda-\sigma-1})(L^2+Q^2)\right]<\frac{\epsilon}{3}. \label{Mcase2} \end{eqnarray} Redefining the operator $h$ for $u\in Y^+$ and $x\in[0,a]$ by \begin{eqnarray*} h(v)(x)=\int\limits_0^\eta\phi(x)K(x,y)(xy)^\sigma v(y)\,dy. \end{eqnarray*} We can now follow the lines of the of the proof of the $W_2$ case in Stewart \cite[Lemma 4.1]{Stewart} to get the proof of case $i=2$. Case $i=3,4$ can be proved analogously as in Giri et al.\ \cite[Lemma 2.2]{AnkikWarnecke}. Then the proof of Lemma \ref{Flemaconvergence} is complete. \hfill{$\Box$} \subsection{The existence result} \begin{thm} Suppose that \textbf{(H1)}-\textbf{(H6)} hold and assume that $u_0\in Y^+$. Then (\ref{elgranasterisco}) has a solution $u\in C\left([0,T],L^1\big(]0,\infty[\big)\right)$. Moreover, we also obtain $u\in C^1\left([0,T],L^1\big(]0,\infty[\big)\right)$ and therefore $u$ is a \emph{regular} solution satisfying (\ref{Fproblema}). \label{Fexisten} \end{thm} \textbf{Proof.} Choose $T,m > 0$, and let $\big(u^n\big)_{n\in\mathbb N}$ be the weakly convergent subsequence of approximating solutions obtained in Lemma \ref{Funlema}. For $t\in[0,T]$ we obtain by weak convergence and Lemma \ref{Flemaproperty}\textbf{(i)} \begin{eqnarray*} \int\limits_0^mxu(x,t)\,dx=\lim_{n\rightarrow\infty}\int\limits_0^mxu^n(x,t)\,dx\leq L(T)<\infty, \end{eqnarray*} and \begin{eqnarray*} \int\limits_{1/m}^mx^{-\sigma}u(x,t)\,dx=\lim_{n\rightarrow\infty}\int\limits_{1/m}^mx^{-\sigma}u^n(x,t)\,dx \leq L(T)<\infty. \end{eqnarray*} Then taking $m\rightarrow\infty$ implies that $u\in Y^+$ with $\|u\|_Y\leq 2L(T)$. Let $\phi\in L^\infty\big(]0,a[\big)$. From Lemma \ref{Funlema} we have for each $s\in[0,t]$ \begin{eqnarray} u^n(s)\rightharpoonup u(s)\quad\mbox{in}\quad L^1\big(]0,a[\big)\quad\mbox{as}\quad n\rightarrow\infty. \label{F3232} \end{eqnarray} From Lemma \ref{Funlema} and Lemma \ref{Flemaconvergence} for each $s\in[0,t]$ we have \begin{eqnarray} \int\limits_0^a\phi(x)\left[M^n(u^n(s))(x)-M(u(s))(x)\right]\,dx\rightarrow 0\quad\mbox{as}\quad n\rightarrow\infty. \label{F341} \end{eqnarray} Also, for $s\in[0,t]$, using Lemma \ref{Flemaproperty}\textbf{(i)} and $\|u\|_Y\leq 2L(T)$ we find that \begin{eqnarray*} \int\limits_0^a\left|\phi(x)\right|\left|M^n(u^n(s))(x)-M(u(s))(x)\right|dx\leq 3\|\phi\|_{L^\infty(]0,a[)}\left[10L(T)+\left(N+1\right)\right]L(T). \end{eqnarray*} Then, by (\ref{F341}), the dominated convergence theorem, and Fubini's Theorem we get \begin{eqnarray} \int\limits_0^t M^n(u^n(s))(x)\,ds\rightharpoonup\int\limits_0^tM(u(s))(x)\,ds\quad\mbox{in}\quad L^1\big(]0,a[\big)\quad\mbox{as}\quad n\rightarrow\infty. \label{F371} \end{eqnarray} From the definition of $M^n$ for $t\in[0,T]$ \begin{eqnarray*} u^n(t)=\int\limits_0^t M^n(u^n(s))\,ds + u^n(0), \end{eqnarray*} and thus it follows by (\ref{F371}), (\ref{F3232}) and the uniqueness of weak limits that \begin{eqnarray} u(t)=\int\limits_0^t M(u(s))\,ds + u(0). \label{388} \end{eqnarray} It follows from the fact that $a$ is arbitrary that $u$ is a solution to (\ref{Fproblema}) on $C\big([0,T],\Omega\big)$. Now we show that $u\in C\left([0,T];L^1\big(]0,\infty[\big)\right)$. Considering $t_n>t$ and by using (\ref{388}) we have that \begin{eqnarray*} \int\limits_0^\infty\left|u(x,t_n)-u(x,t)\right|dx\!\!&\!\!=\!\!&\!\!\int\limits_0^\infty\left|\frac{1}{2}\int\limits_t^{t_n}\int\limits_0^xK(x-y,y)u(x-y,\tau)u(y,\tau)\,dy\,d\tau\right.\\ &&\qquad\qquad-\int\limits_t^{t_n}\int\limits_0^\infty K(x,y)u(x,\tau)u(y,\tau)\,dy\,d\tau\\ &&\qquad\qquad\left.+\int\limits_t^{t_n}\int\limits_x^\infty b(x,y)S(y)u(y,\tau)\,dy\,d\tau-\int\limits_t^{t_n}S(x)u(x,\tau)\,d\tau\right|dx\\ &\leq&\int\limits_t^{t_n}\left[\frac{3}{2}\int\limits_0^\infty\int\limits_0^\infty K(x,y)u(x,\tau)u(y,\tau)\,dy\,dx\right.\\ &&\qquad\qquad\left.+\int\limits_0^\infty\int\limits_0^yb(x,y)S(y)u(y,\tau)\,dx\,dy+\int\limits_0^\infty S(x)u(x,\tau)\,dx\right]d\tau. \end{eqnarray*} By using the definition (\ref{FC1}) of $C_1$, Lemma \ref{Flemaproperty} \textbf{(i)}, \textbf{(H3)}, \textbf{(H4)} and (\ref{FbN}) we find that \begin{eqnarray} \int\limits_0^\infty\left|u(x,t_n)-u(x,t)\right|dx\!\!&\!\!\leq\!\!&\!\!\int\limits_t^{t_n}\left[\frac{3}{2}\int\limits_0^\infty\int\limits_0^\infty (1+x)^\lambda(1+y)^\lambda(xy)^{-\sigma}u(x,\tau)u(y,\tau)\,dy\,dx\right.\nonumber\\ &&\qquad\qquad+\left.N\int\limits_0^\infty yu(y,\tau)\,dy+\int\limits_0^\infty xu(x,\tau)\,dx\right]d\tau\nonumber\\ &\leq&\int\limits_t^{t_n}\left[\frac{3}{2}C_1^2\int\limits_0^\infty\int\limits_0^\infty\Big((xy)^{-\sigma}+x^{\lambda-\sigma}y^{-\sigma}+y^{\lambda-\sigma}x^{-\sigma}\Big)u(x,\tau)u(y,\tau)\,dy\,dx\right.\nonumber\\ &&\qquad\qquad+\left.N\int\limits_0^\infty yu(y,\tau)\,dy+\int\limits_0^\infty xu(x,\tau)\,dx\right]d\tau\nonumber\\ &\leq&\left[\frac{9}{2}C_1^2L^2(T)+(N+1)L(T)\right](t_n-t). \label{389} \end{eqnarray} Then from (\ref{389}) we obtain that \begin{eqnarray} \int\limits_0^\infty\left|u(x,t_n)-u(x,t)\right|dx\rightarrow 0\quad\mbox{as}\quad t_n\rightarrow t. \label{390} \end{eqnarray} The same argument holds when $t_n<t$. Hence (\ref{390}) holds for $\left|t_n-t\right|\rightarrow 0$ and we can conclude that $u\in C\left([0,T];L^1\big([0,\infty[\big)\right)$. Now, we have that our solution satisfies the integral equation \begin{eqnarray} u(x,t)\!\!&\!\!=\!\!&\!\!u(x,0)+\int\limits_0^t\left[\frac{1}{2}\int\limits_0^xK(x-y,y)u(x-y,\tau)u(y,\tau)\,dy-\int\limits_0^\infty K(x,y)u(x,\tau)u(y,\tau)\,dy\right. \nonumber\\ \!\!&\!\!\!\!&\!\!\qquad\qquad\qquad\left.+\int\limits_x^\infty b(x,y)S(y)u(y,t)\,dy-S(x)u(x,t)\right]d\tau.\label{36} \end{eqnarray} From this we can see that for $u$, which is a continuous function in time $t$, that the integrand \begin{eqnarray} f(x,t)\!\!&\!\!=\!\!&\!\!\frac{1}{2}\int\limits_0^xK(x-y,y)u(x-y,\tau)u(y,\tau)\,dy-\int\limits_0^\infty K(x,y)u(x,\tau)u(y,\tau)\,dy \nonumber\\ \!\!&\!\!\!\!&\!\!\qquad+\int\limits_x^\infty b(x,y)S(y)u(y,t)\,dy-S(x)u(x,t)\label{37} \end{eqnarray} is also a continuous function in time. We now show that $f(\cdot,t)\in L^1\big([0,\infty[\big)$ for any $t\in[0,T]$. Integrating (\ref{37}) from $0$ to $\infty$ w.r.t. $x$ we have to show that the following integral is bounded \begin{eqnarray} \int\limits_0^\infty f(x,t)dx\!\!&\!\!=\!\!&\!\!\frac{1}{2}\int\limits_0^\infty\int\limits_0^xK(x-y,y)u(x-y,\tau)u(y,\tau)\,dy\,dx -\int\limits_0^\infty\int\limits_0^\infty K(x,y)u(x,\tau)u(y,\tau)\,dy\,dx \nonumber\\ \!\!&\!\!\!\!&\!\!\qquad+\int\limits_0^\infty\int\limits_x^\infty b(x,y)S(y)u(y,t)\,dy\,dx-\int\limits_0^\infty S(x)u(x,t)\,dx.\label{38} \end{eqnarray} Working with the second, third and fourth terms of the right hand side of (\ref{38}) as in (\ref{389}) we find that \begin{eqnarray} &&\int\limits_0^\infty\int\limits_0^\infty K(x,y)u(x,\tau)u(y,\tau)\,dy\,dx<\infty,\label{39-1} \\ &&\int\limits_0^\infty\int\limits_x^\infty b(x,y)S(y)u(y,t)\,dy\,dx=\int\limits_0^\infty\int\limits_0^y b(x,y)S(y)u(y,t)\,dx\,dy<\infty\quad\mbox{and},\label{39-2}\\ &&\int\limits_0^\infty S(x)u(x,t)\,dx<\infty\label{39-3} \end{eqnarray} Now, by Tonelli's Theorem \cite[page 293]{Nielsen} we have that \begin{eqnarray*} \int\limits_0^\infty\int\limits_0^xK(x-y,y)u(x-y,\tau)u(y,\tau)\,dy\,dx=\int\limits_0^\infty\int\limits_y^\infty K(x-y,y)u(x-y,\tau)u(y,\tau)\,dy\,dx \end{eqnarray*} holds if \begin{eqnarray*} \int\limits_0^\infty\int\limits_0^xK(x-y,y)u(x-y,\tau)u(y,\tau)\,dy\,dx<\infty \end{eqnarray*} or \begin{eqnarray*} \int\limits_0^\infty\int\limits_y^\infty K(x-y,y)u(x-y,\tau)u(y,\tau)\,dy\,dx<\infty. \end{eqnarray*} Making a change of varible $x-y = x'$, $y = y'$ in the second integral term we find, by using the symmetry of $K(x, y)$, that using (\ref{39-1}) \begin{eqnarray*} \int\limits_0^\infty\int\limits_y^\infty K(x-y,y)u(x-y,\tau)u(y,\tau)\,dx\,dy\!\!&\!\!=\!\!&\!\!\int\limits_0^\infty\int\limits_0^\infty K(x',y')u(x',\tau)u(y',\tau)\,dx'\,dy'\nonumber\\ &=&\int\limits_0^\infty\int\limits_0^\infty K(x',y')u(x',\tau)u(y',\tau)\,dy'\,dx'<\infty. \end{eqnarray*} From this it follows that \begin{eqnarray} \int\limits_0^\infty\int\limits_0^x K(x-y,y)u(x-y,\tau)u(y,\tau)\,dx\,dy\!\!&\!\!=\!\!&\!\!\int\limits_0^\infty\int\limits_0^\infty K(x',y')u(x',\tau)u(y',\tau)\,dy'\,dx'<\infty. \label{40} \end{eqnarray} Then, from (\ref{39-1})-(\ref{39-3}) and (\ref{40}) together with (\ref{38}) it follows that $f(\cdot,t)\in L^1\big([0,\infty[\big)$. Moreover, we have that $f\in C\left([0,T];L^1\big([0,\infty[\big)\right)$. Then, using this fact, (\ref{36}), (\ref{37}), and $u(x,0)\in Y^+$ we find that \begin{eqnarray*} u(x,t)\!\!&\!\!=\!\!&\!\!u(x,0)+\int\limits_0^t\left[\frac{1}{2}\int\limits_0^xK(x-y,y)u(x-y,\tau)u(y,\tau)\,dy-\int\limits_0^\infty K(x,y)u(x,\tau)u(y,\tau)\,dy\right. \nonumber\\ \!\!&\!\!\!\!&\!\!\qquad\qquad\qquad\left.+\int\limits_x^\infty b(x,y)S(y)u(y,t)\,dy-S(x)u(x,t)\right]d\tau. \end{eqnarray*} gives $u\in C^1\left([0,T];L^1\big(]0,\infty[\big)\right)$ since the right hand side lies in this space. And this completes the proof of Theorem \ref{Fexisten}.\hfill{$\Box$} \section{Uniqueness of Solutions} \begin{thm} If \textbf{(H1)}, \textbf{(H2)}, \textbf{(H3')}, \textbf{(H4')}, and \textbf{(H5)} hold then the problem (\ref{Fproblema})-(\ref{Fcondinicial}) has a unique solution $u\in C\left([0,T],L^1\big(]0,\infty[\big)\right)$. \end{thm} \textbf{Proof}: Let us consider $u_1$ and $u_2$ to be solutions to (\ref{Fproblema})-(\ref{Fcondinicial}) on $[0,T]$ for $T>0$ arbitrarily chosen, with $u_1(x,0)=u_2(x,0)$ and set $U=u_1-u_2$. We define for $n=1,2,3,\ldots$ \begin{eqnarray*} m^n(t)=\int\limits_0^n (x^{-\sigma}+x^{\lambda-\sigma})\left|U(x,t)\right|dx. \end{eqnarray*} Then, proceeding analogously as in \cite{Stewartuniq} we have that \begin{eqnarray*} u_1(x,t)=u_2(x,t)\quad\mbox{for a.e.}\quad x\in]0,\infty[. \end{eqnarray*} For details see Cueto Camejo \cite{CuetoCamejo}. \hfill{$\Box$} \phantomsection \section*{Acknowledgements} This work was supported by the International Max-Planck Research School, `Analysis, Design and Optimization in Chemical and Biochemical Process Engineering', Otto-von-Guericke-Universit\"at Magdeburg. The authors thank gratefully for the funding of C. Cueto Camejo through this PhD program by the state of Saxony-Anhalt.
\section*{Supplementary information for ``Violation of Entropic Leggett-Garg Inequality in Nuclear Spins"} \title{Violation of Entropic Leggett-Garg Inequality in Nuclear Spins} \author{Hemant Katiyar$^1$, Abhishek Shukla$^1$, Rama Koteswara Rao$^2$, and T. S. Mahesh$^1$} \noindent\textbf{Encoding the probability using a CNOT gate:} Consider a system qubit initially prepared in a general state $\rho_S$ and an ancilla qubit prepared in the state $\ket{0}\bra{0}$. The state of the system qubit can be written as $P(0_i) \ket{0}\bra{0} + P(1_i) \ket{1}\bra{1} + a \ket{1}\bra{0} + a^\dagger \ket{0}\bra{1}$, where $a$ is a complex probability amplitude. The CNOT gate encodes the probability of the outcomes in the diagonal elements of ancilla qubit since, \begin{eqnarray} ( P(0_i) \ket{0}\bra{0}_S + P(1_i) \ket{1}\bra{1}_S + a \ket{1}\bra{0}_S + a^\dagger \ket{0}\bra{1}_S ) \otimes \ket{0}\bra{0}_A \stackrel{\mathrm{CNOT}}{\longrightarrow} & \ket{0}\bra{0}_S \otimes P(0_i) \ket{0}\bra{0}_A + \ket{1}\bra{1}_S \otimes P(1_i) \ket{1}\bra{1}_A & \nonumber \\ &+ \ket{1}\bra{0}_S \otimes a \ket{1}\bra{0}_A + \ket{0}\bra{1}_S \otimes a^\dagger \ket{0}\bra{1}_A. & \nonumber \end{eqnarray} The probabilities $P(0_i)$ and $P(1_i)$ can now be retrieved by tracing over the system qubit and reading the diagonal elements of the ancilla state.\\ \noindent\textbf{The qubit systems:} Fig. \ref{tfie} shows the molecular structures and the Hamiltonian parameters of $^{13}$CHCl$_3$ (Fig. \ref{tfie}(a,b)) and trifluoroiodoethylene (Fig. \ref{tfie}(c,d)). In the case of $^{13}$CHCl$_3$, spin-lattice (T1) and spin-spin (T2) relaxation time constants for the $^{1}$H spin are, respectively, 4.1 s and 4.0 s. The corresponding time constants for $^{13}$C are 5.5 s and 0.8 s. In the case of trifluoroiodoethylene, the effective $^{19}$F transverse relaxation time constants ($T_2^*$) were about 0.8 s and their longitudinal relaxation time constants were all longer than 6.3 s.\\ \begin{figure}[H] \centering \vspace*{-0.5cm} \includegraphics[width=6.8cm,height=10cm,angle=90]{exptscheme1.eps} \vspace*{-3cm} \caption{The molecular structures of chloroform (a) and trifluoroiodoethylene (c) and the corresponding tables (b and d) of relative resonance frequencies (diagonal elements) and the J-coupling constants. The pulse sequence for initializing trifluoroiodoethylene is shown in (e). In (e) the open pulses are $\pi$ pulses and the delay $\tau = 1/(4J_{23})$.} \label{tfie} \end{figure} \noindent\textbf{Initialization in two-qubit ($^{13}$CHCl$_3$) system:} The initialization involved preparing the maximally mixed state $\rho_S = \mathbbm{1}/2$ on the system qubit ($^{13}$C). This is achieved by a $\pi/2$ pulse on $^{13}$C followed by a strong Pulsed Field Gradient (PFG).\\ \noindent\textbf{Initialization in three-qubit (trifluoroiodoethylene) system:} The equilibrium deviation density matrix evolves under the PPS sequence (Fig. \ref{tfie}(e)) as follows \begin{eqnarray} S_{1z} + S_{2z} + S_{3z} \stackrel{(\pi/2)_{1x} (\pi/3)_{3x}, \mathrm{PFG}}{\longrightarrow} S_z^2 + \frac{1}{2}S_{3z} \stackrel{(\pi/4)_{2x}}{\longrightarrow} \frac{1}{\sqrt{2}}S_{2z} - \frac{1}{\sqrt{2}}S_{2y} + \frac{1}{2}S_{3z} \stackrel{1/(2J_{23})}{\longrightarrow} &\frac{1}{\sqrt{2}}S_{2z} + \sqrt{2}S_{2x} S_{3z} + \frac{1}{2}S_{3z}& \nonumber \\ &\downarrow& \hspace*{-1.5cm}(\pi/4)_{-2y},\mathrm{PFG} \nonumber \\ &\frac{1}{2} (S_{2z} + 2S_{2z} S_{3z} + S_{3z} ). \nonumber& \end{eqnarray} The above deviation density matrix is equivalent to the traceless part of $\frac{1-\epsilon}{8}\mathbbm{1} + \epsilon \left\{ \frac{1}{2}\mathbbm{1}_S \otimes \ket{00}\bra{00}_A \right\}$ where $\epsilon \sim 10^{-5}$ is the purity factor \cite{cory}.\\ \noindent\textbf{Diagonal tomography:}\\ The diagonal tomography was carried out using a strong PFG to dephase out all the residual off-diagonal elements and using a $6^\circ$ non-selective linear readout pulse. The resulting intensities of the spectral lines constrain the diagonal elements ($d_i$) of the traceless deviation density matrix. The experimental deviation density matrix is normalized w.r.t. the theoretical traceless density matrix such that they both have the same root mean square value $\sqrt{\sum_i d_i^2}$, and trace is introduced by adding identity matrix to the normalized deviation matrix. The resulting diagonal density matrix yields the probabilities. Estimation of random errors were carried out by several repetitions at each $\theta$ value.\\ \end{titlepage} \end{document}
\section{Distant graph and connected components} The \emph{projective line} $\bP(R)$ over any ring $R$ (associative with $1\neq 0$) can be defined in terms of the free left $R$-module $R^2$ as follows \cite{blunck+he-05}, \cite{herz-95a}: It is the orbit of a starter point $R(1,0)$ under the action of the general linear group $\GL_2(R)$ on $R^2$. A basic notion on $\bP(R)$ is its \emph{distant relation}: Two points are called distant (in symbols: $\dis$) if they can be represented by the elements of a two-element basis of $R^2$. The \emph{distant graph} $(\bP(R),\dis)$ has as vertices the points of $\bP(R)$ and as edges the pairs of distant points. The distant graph is connected precisely when $\GL_2(R)$ is generated by the \emph{elementary linear group} $\E_2(R)$, i.e., the subgroup of $\GL_2(R)$ which is generated by elementary transvections, together with the set of all invertible diagonal matrices \cite{blunck+h-01a}. The orbit of $R(1,0)$ under $\E_2(R)$ is a connected component of the distant graph. It admits a parametrisation in terms of infinitely many formulas \cite{blunck+h-01a}, \cite{blunck+h-03b}. The situation is less intricate for a ring $R$ of \emph{stable rank} $2$ (see \cite{chen-11a}, \cite{veld-85}, or \cite{veld-95}), as it gives rise to a connected distant graph with diameter $\leq 2$. The above-mentioned parametrisation turns into \emph{Bartolone's parametrisation} \cite{bart-89} of $\bP(R)$, namely \begin{displaymath} \bP(R) = \{ R(t_2t_1-1,t_2)\mid t_1,t_2\in R \} \mbox{~~~~~($R$ of stable rank $2$)} . \end{displaymath} Refer to the seminal paper of P.~M.~Cohn \cite{cohn-66} for the algebraic background, and to the work of A.~Blunck \cite{blunck-97a}, \cite{blunck-02a} for orbits of the point $R(1,0)$ under other subgroups of $\GL_2(R)$. \section{Chain Geometries, subspaces and Jordan Systems} Let $R$ be an algebra over a commutative field $K$; by identifying $K$ with $K\cdot 1_R$ the projective line $\bP(K)$ is embedded in $\bP(R)$. For $R\neq K$ the projective line $\bP(R)$ can be considered as the point set of the \emph{chain geometry} $\Sigma(K,R)$; the $\GL_2(R)$ orbit of $\bP(K)$ is the set of \emph{chains} \cite{blunck+he-05}, \cite{herz-95a}. The geometries of M\"{o}bius, Minkowski and Laguerre are well known examples of chain geometries \cite{benz-73}. A crucial property is that any three mutually distinct points are on a unique chain. The chain geometry $\Sigma(K,R)$ may be viewed as a refinement of the distant graph, since two points of $\bP(R)$ are distant if, and only if, they are on a common chain. There are cases though, when the word ``refinement'' is inappropriate in its strict sense: Let $R=\End_F(V)$ be the endomorphism ring of a vector space $V$ over a (not necessarily commutative) field $F$ and let $K$ denote the \emph{centre} of $F$. Then the $K$-chains of $\bP(R)$ can be defined solely in terms of the distant graph $(\bP(R),\dis)$ \cite{blunck+h-12z}. Each chain geometry $\Sigma(K,R)$ is a \emph{chain space}; see \cite{blunck+he-05}, where also the precise definition of \emph{subspaces} of a chain space is given. The algebraic description of subspaces of $\Sigma(K,R)$ is due to A.~Herzer \cite{herz-92b} and H.-J.~Kroll \cite{kroll-91a}, \cite{kroll-92b}, \cite{kroll-92a}. It is based on the following notions: A \emph{Jordan system} is a $K$-subspace of $R$ satisfying two extra conditions: (i) $1\in J$; (ii) If $b\in J$ has an inverse in $R$ then $b^{-1}\in J$. (See \cite{loos-75} for relations with \emph{Jordan algebras} and \emph{Jordan pairs} and compare with \cite{gold-06a}, \cite{matt-07a}.) A Jordan system $J$ is called \emph{strong} if it satisfies a (somewhat technical) condition which guarantees the existence of ``many'' invertible elements in $J$. Strong Jordan systems are closed under \emph{triple multiplication}, i.~e., $ xyx\in J$ for all $x,y\in J$. The \emph{projective line $\bP(J)$ over a strong Jordan system} $J\subset R$ is defined by restricting the \emph{parameters} $t_1,t_2$ to $J$ in Bartolone's parametrisation. We wish to emphasise that in general a point of $\bP(J)$ cannot be written as $R(a,b)$ with $a,b\in J$, unless $J$ is even a subalgebra of $R$. The main theorem about subspaces is as follows: If $R$ is a strong algebra then any connected subspace of $\Sigma(K,R)$ is projectively equivalent to a projective line over a strong Jordan system of $R$. Projective lines over strong Jordan systems admit many applications: For example, one may use them to describe subsets of Grassmannians which are closed under reguli \cite{herz-92b} or chain spaces on quadrics \cite{blunck-97}. See also \cite{blunck-94}, \cite{herz-08a}, \cite{herz-10a}, \cite{herz-11a}, and the numerous examples given in \cite{blunck+he-05}. \par Finally, let us mention one of the many questions that remain: \emph{Is it possible to replace the strongness condition for Jordan systems by closedness under triple multiplication without affecting the known results?} A partial affirmative answer was given in \cite{blunck+h-10a} for case when $R$ is the ring of $n\times n$ matrices over a field $F$ with an involution $\sigma$ and $J$ is the (not necessarily strong) Jordan system of $\sigma$-Hermitian matrices. The proof is based upon the verification that the projective line over this $J$ is, up to some notational differences, nothing but the point set of a \emph{dual polar space} \cite{cameron-82a} or, in the terminology of \cite{wan-96}, the point set of a \emph{projective space of $\sigma$-Hermitian matrices}. \par A wealth of further references can be found in \cite{benz-73}, \cite{blunck+he-05}, \cite{havl-07a}, \cite{herz-95a}, \cite{huanglp-06a}, \cite{pank-10a}, \cite{veld-95}, and \cite{wan-96}. Refer to \cite{brehm-08}, \cite{brehm+g+s-95}, \cite{faure-04a}, \cite{havl+m+p-11a}, \cite{havl+k+o-12z}, \cite{havlicek+saniga-09a}, and \cite{lash-97} for deviating definitions of projective lines which we cannot present here.
\section{Introduction} Suppose that an area is being monitored by a number of sensors which transmit their observations to a central location, that we will call fusion center. At some unknown time, an abrupt disorder occurs, such as an unexpected intrusion, and changes the dynamics of the observed processes in all sensors simultaneously. The goal is to raise an alarm at the fusion center as soon as possible after the occurrence of the change. When the sensors transmit their complete observations to the fusion center, this is the classical problem of sequential change detection, for exhaustive reviews on which we refer to \cite{niki}, \cite{poor}, \cite{lairoyal}, \cite{shir}, \cite{polu}. However, classical detection rules typically are not applicable in modern application areas, such as mobile and wireless communications and distributed surveillance systems. In such systems, the sensors are typically low-power devices whose links with the fusion center are characterized by limited communication bandwidth \cite{vij},\cite{veera}. Thus, in order to preserve the robustness of the network, it is necessary to limit the overall communication load and, in particular, the transmission activity of each sensor. This primarily implies a \textit{quantization} constraint, i.e., each sensor should transmit a small number of bits each time it communicates with the fusion center, but also a \textit{rate} constraint, i.e., each sensor should communicate with the fusion center at a lower rate than its sampling rate. As a result, before constructing a sequential detection rule at the fusion center, the designer must first decide what kind of information should be transmitted from the sensors, taking into account the above communication constraints. In what follows, we will call detection rules that respect such constraints \textit{decentralized}, in contrast to the \textit{centralized} ones that require knowledge of the full sensor observations. Most papers in the decentralized literature (see, e.g., \cite{crow}, \cite{veer}, \cite{veera}, \cite{tart}) assume that each sensor transmits a quantized version of \textit{every} observation it takes, i.e., the communication rate is equal to the sampling rate. For a discussion on one-shot schemes, where each sensor transmits to the fusion center a single bit \textit{at most once}, we refer to \cite{moustdec}. A decentralized detection rule which enjoys an asymptotic optimality property was proposed by Mei \cite{mei}, however the performance of this scheme in practice is often worse than that of asymptotically suboptimal detection rules. Thus, it has been an open problem to find an asymptotically optimal decentralized detection rule that is also efficient in practice. The main contribution of this work is that we propose such a rule. Specifically, we suggest that each sensor communicates with the fusion center at stopping times of its local filtration; at every communication, it transmits a \textit{low-bit} message which ``summarizes'' the evolution of its local sufficient statistic since the previous communication; the fusion center, in parallel, runs a CUSUM test on the transmitted messages in order to detect the change. For similar communication schemes in the context of decentralized sequential hypothesis testing we refer to \cite{fel} and \cite{yil}. The design and analysis of the proposed scheme, that we call D-CUSUM, is different in discrete and continuous time. However, in both cases we establish a \textit{second-order} asymptotic optimality property, that is stronger than the first-order asymptotic optimality of the detection rule in \cite{mei}. In particular, we show that the performance loss of D-CUSUM with respect to the optimal centralized CUSUM remains bounded as the period of false alarms goes to infinity. Moreover, we show that D-CUSUM remains first-order asymptotically optimal even when it induces an asymptotically low communication rate and there is an asymptotically large number of sensors. Simulation experiments suggest that these strong theoretical properties are also accompanied by very good performance in practice and that D-CUSUM is much more efficient than a similar, CUSUM-based decentralized detection rule that relies on communication from the sensors at deterministic times. In what follows, in Section 2, we formulate the problem of (decentralized) sequential change detection and describe the main decentralized schemes in the literature. In Section 3, we define and analyze the proposed scheme both in continuous and in discrete time. In Section 4, we summarize and discuss an extension in the case of correlated sensors. The proof of all results, as well as some supporting lemmas, are presented in Appendices A-E. \section{Sequential Change Detection} Let $\{(\xi_{t}:=\xi_{t}^{1},\ldots, \xi_{t}^{K})\}$ be a $K$-dimensional stochastic process, where $\xi_{0}^{k}:=0$ and $\xi^{k}$ is the observed process at sensor $k$, $1 \leq k \leq K$. We denote by $\{ {\cF}_{t}^{k}\}$ the local filtration at sensor $k$ and by $\{ {\cF}_{t}\}$ the global filtration, i.e., $ {\cF}_{t}^{k}:= \sigma(\xi_{s}^k, \, 0 \leq s \leq t)$ and $\mathscr{F}_{t}:= \vee_{k} \mathscr{F}_{t}^{k}$. Time may be either discrete $(t \in \mathbb{N})$ or continuous $(t \in [0, \infty))$ and in the latter case all filtrations are considered to be right-continuous. We assume that at some unknown, deterministic time $\tau \geq 0$, the distribution of $\xi$, which we denote by $\mathsf{P}_\tau$, changes from $\mathsf{P}_{\infty}$ to $\mathsf{P}_{0}$, where $\mathsf{P}_{0}$ and $\mathsf{P}_{\infty}$ are two completely specified, locally equivalent probability measures on the canonical space of $\xi$. In other words, $\mathsf{P}_\tau$ coincides with $\mathsf{P}_{\infty}$ when both measures are restricted to $ {\cF}_{t}$ and $t \leq \tau$, whereas for $t> \tau$ we can define the following log-likelihood ratio process $ u_{t}-u_{\tau}:=\log \frac{\text{d} \mathsf{P}_{\tau}}{\text{d} \mathsf{P}_{\infty}} \Big|_{ {\cF}_{t}} , \quad t \geq \tau; \quad u_{0}:=0. $ \subsection{The centralized setup} In the centralized setup, where the fusion center has access to all sensor observations, the problem is to find an $\{ {\cF}_{t}\}$-stopping time $\mathcal{T}$ that has small detection delay and rare false alarms, i.e., $\mathcal{T}$ should take large values under $\mathsf{P}_{\infty}$ and $\mathcal{T}-\tau$ small values under $\mathsf{P}_{\tau}$. There are different approaches in how to quantify detection delay and false alarms, such as the Bayesian formulation due to Shiryaev \cite{shiry} (see also \cite{bei2}, \cite{pes}, \cite{gap}, \cite{savas}, \cite{sezer}) or the minimax formulation due to Pollak \cite{poll} (see also \cite{polu}, \cite{tpp}). In this work, we focus on the formulation suggested by Lorden \cite{lord}, where the performance of a detection rule $\mathcal{T}$ is measured by its worst-case (with respect to $\tau$) conditional expected delay given the worst possible history of observations up to $\tau$, \begin{equation} \label{lord} \mathcal{J}_{\text{L}}[\mathcal{T}] = \sup_{\tau \geq 0} \, \text{ess\,sup} \; \mathsf{E}_{\tau} [ (\mathcal{T}-\tau)^{+} | \mathscr{F}_{\tau} ], \end{equation} and an optimal detection rule is a solution to the following optimization problem \begin{equation} \label{lord_crit} \inf_{\mathcal{T}} \mathcal{J}_{\text{L}}[\mathcal{T}] ~ \text{when} ~ \mathsf{E}_{\infty} [\mathcal{T}] \geq \gamma , \end{equation} where $\gamma>0$. In other words, the goal in this approach is to minimize the detection delay under the worst-case scenario with respect to both the changepoint and the history of observations before the change, while controlling the period of false alarms above a desired level, $\gamma$. It is well known (see \cite{moust}, \cite{moustext}) that when $\{u_{t}\}_{t \in \mathbb{N}}$ is a random walk, the solution to this problem is given by Page's \cite{page} Cumulative Sums (CUSUM) test, \begin{equation} \label{cusum} \cus:= \inf \{ t \geq 0: y_{t} \geq \nu \} , ~ \text{where} \quad y_{t}:= u_{t} - \inf_{0 \leq s < t} u_{s}, \end{equation} and $\nu$ is defined so that the false alarm constraint in (\ref{lord_crit}) be satisfied with equality, i.e., $\mathsf{E}_{\infty} [\cus]=\gamma$. This exact (i.e., non-asymptotic) optimality of the CUSUM test can be extended to a much richer class of dynamics if we adopt an idea of Liptser and Shiryaev \cite{lip} and measure detection delay and period of false alarms not in terms of actual time, but in terms of Kullback-Leibler divergence. Indeed, working similarly to \cite{moustito}, we replace the performance measure $\mathcal{J}_{L}$ by \begin{equation} \label{mod} \mathcal{J}[\mathcal{T}] := \sup_{\tau \geq 0} \, \text{ess\,sup} \; \mathsf{E}_{\tau} [ ( u_{\mathcal{T}}- u _{\tau} )\ind{\mathcal{T}>\tau} | \mathscr{F}_{\tau} ] \end{equation} and define an optimal detection rule as a solution to \begin{equation} \label{mod_crit} \inf_{\mathcal{T}} \mathcal{J}[\mathcal{T}] ~ \text{when} ~ \mathsf{E}_{\infty} [-u_{\mathcal{T}}] \geq \gamma, \end{equation} a problem that is equivalent to \eqref{lord_crit} when $\{u_{t}\}$ is a random walk. However, it has been shown in \cite{moustito}, \cite{chro} that the CUSUM test, with threshold $\nu$ chosen so that $\mathsf{E}_{\infty} [-u_{\cus}] = \gamma$, also solves problem \eqref{mod_crit} whenever $\{u_{t}\}$ has continuous paths and \begin{equation} \label{full} \lim_{t\rightarrow \infty} \langle u \rangle_{t}=\infty \quad \mathsf{P}_{0}, \mathsf{P}_{\infty}-\text{a.s.}, \end{equation} where $\langle u \rangle_{t}$ is the quadratic variation of $u_t$. The latter optimality result implies that CUSUM solves Lorden's original problem (\ref{lord_crit}) whenever $\{u_{t}\}$ has continuous paths and $\langle u\rangle_{t}$ is proportional to $t$. This is the case, for example, when each $\xi^{k}$ is a fractional Brownian motion (fBm) with Hurst index $H$ before the change and adopts a polynomial drift term with exponent $H+1/2$ after the change \cite{chro}. In the special case that $H=1/2$, this implies the well-known optimality of CUSUM for detecting a constant drift in a Brownian motion, established by Shiryaev \cite{shircus} and Beibel \cite{bei}. \subsection{The decentralized setup} Centralized (classical) detection rules as the CUSUM test cannot be applied in a decentralized setup, where communication constraints must be taken into account. In this context, before defining a detection rule at the fusion center, we must first specify a \textit{communication scheme}, that will determine the information that will be transmitted from the sensors to the fusion center. Therefore, we define a \textit{decentralized} sequential detection rule as a pair $(\{\tilde{ {\cF}_{t}}\}, \mathcal{T})$, where $\mathcal{T}$ is an $\{\tilde{ {\cF}_{t}}\}$-stopping time and $\{\tilde{ {\cF}_{t}}\}$ is a filtration of the form \begin{equation} \label{flow} \tilde{ {\cF}_{t}}:= \sigma (( \tau^k_{n},z_{n}^k): \tau_{n}^k \leq t, k=1, \ldots, K), \end{equation} where each $\{\tau_{n}^k\}_{n \in \mathbb{N}}$ is the sequence of communication times for sensor $k$ and $z_{n}^k$ is the message transmitted to the fusion center at time $\tau_{n}^{k}$. Each $\tau_{n}^k$ must be an $\{ {\cF}_{t}^k\}$-stopping time and each $z_{n}^k$ an $\mathscr{F}^k_{\tau_{n}^k}$-measurable random variable that takes values in a \textit{finite} set, so that a small number of bits is required for its transmission to the fusion center. Moreover, since many applications are characterized by limited storage capacity, we require additionally that each $z_{n}^{k}$ is measurable with respect to $\sigma(\xi_{s}^{k}, \; \tau_{n-1}^{k} \leq s \leq \tau_{n}^{k})$, the $\sigma$-algebra generated by the observations at sensor $k$ between its $n-1$ and $nth$ transmission. Note that this framework forbids communication between sensors or feedback from the fusion center to the sensors. Such possibilities impose a much heavier communication load on the network and raise questions regarding the design of the network architecture, which we do not consider here. For decentralized detection rules that require feedback we refer to \cite{veer}. Ideally, we would like to find the best possible decentralized detection rule, performing a joint optimization over the communication scheme at the sensors and the detection rule at the fusion center. Such an optimization problem is highly intractable, even if one makes a number of simplifying assumptions \cite{veer}. For this reason, we will use the centralized CUSUM as the ultimate benchmark and compare any decentralized detection rule against it. We can only hope that such a detection rule attains the optimal centralized performance asymptotically. Thus, if $(\{\tilde{ {\cF}_{t}}\},\mathcal{T})$ is an arbitrary decentralized detection rule and $\mathcal{S}$ the centralized CUSUM test so that $\mathsf{E}_{\infty}[-u_{\mathcal{T}}]\ge\gamma= \mathsf{E}_{\infty}[-u_{\mathcal{S}}]$ for any $\gamma>0$, we will say that $\mathcal{T}$ is asymptotically optimal of \textit{first} order if $\mathcal{J}[\mathcal{T}]/\mathcal{J}[\mathcal{S}] \rightarrow 1$ as $\gamma \rightarrow \infty$ and of \textit{second} order if $\mathcal{J}[\mathcal{T}] - \mathcal{J}[\mathcal{S}] = \mathcal{O}(1)$ as $\gamma\to\infty$. Clearly, since $\mathcal{J}[S] \rightarrow \infty$ as $\gamma \rightarrow \infty$, second order asymptotic optimality is a stronger property, which guarantees that the inflicted performance loss remains bounded as the rate of false alarms goes to 0. As it is common in the literature of decentralized sequential detection, we will assume that observations from different sensors are independent. Thus, if $\mathsf{P}_{\tau}^{k}$ is the distribution of $\xi^{k}$, then $\mathsf{P}_{\tau}= \mathsf{P}_{\tau}^{1} \times \ldots \times \mathsf{P}_{\tau}^{K}$ for any $\tau \in [0, \infty]$ and, consequently, $ u_{t}:= u_{t}^{1}+ \ldots + u_{t}^{K}, \quad \text{where} \quad u_{t}^k :=\log \frac{\text{d} \mathsf{P}_{0}^{k}}{\text{d} \mathsf{P}_{\infty}^{k}} \Big|_{ {\cF}_{t}^k}, $$ for any $t \geq 0$. We also assume that the local Kullback-Leibler (KL) information numbers, $I_{0}^k:= \mathsf{E}_{0}[u_{1}^k]$ and $I_{\infty}^k:= -\mathsf{E}_{\infty}[u_{1}^k]$, are positive and finite for every $1\leq k \leq K$ and, furthermore, we define the corresponding average KL-numbers \begin{equation} \label{klaver} \bar{I}_{0}:= \frac{1}{K}\mathsf{E}_{0}[u_{1}]= \frac{1}{K}\sum_{k=1}^{K} I_{0}^k \quad \text{and} \quad \bar{I}_{\infty}:= \frac{1}{K}\mathsf{E}_{\infty}[-u_{1}]=\frac{1}{K}\sum_{k=1}^{K} I_{\infty}^k. \end{equation} In the remainder of this section, we describe the main decentralized sequential detection rules in the literature, embedding them in the above framework. We classify them into two categories; in the first, the sensors transmit systematically compressed versions of their data to the fusion center and the latter combines the received messages in order to detect the change; in the second, each sensor detects individually the change and the fusion center combines the local sensor decisions. \subsubsection{Q-CUSUM} Suppose that each sensor transmits to the fusion center quantized versions of its local log-likelihood ratio process at deterministic, equidistant times. Specifically, if for each sensor the communication period is $r$ and the available alphabet $\{1,\ldots, b\}$, where $b\geq 2$ is an integer, then \begin{equation} \label{zdd} \tau_{n}^k=rn \; \text{and} \; z_n^k= \sum_{j=1}^{b} j \, \ind{\Gamma_{j-1}^k \leq u^k_{rn}- u^k_{r(n-1)} < {\Gamma}_{j}^k}, \end{equation} where $-\infty=:\Gamma_{0}^k< {\Gamma}_{1}^k<\ldots < {\Gamma}_{b}^k:=\infty$ are fixed thresholds. This communication scheme induces synchronous communication to the fusion center, which receives at each time $\tau_{n}^k=rn$ the $K$-dimensional vector $(z_{n}^{1}, \ldots, z_{n}^{K})$. If we additionally assume that each $\{u_{t}^k\}$ has stationary and independent increments, then a natural detection rule at the fusion center is the corresponding CUSUM stopping time \begin{equation} \label{eq:cus2} \hat{\mathcal{S}} := r \cdot \inf\{n \in \mathbb{N} : \hat{y}_{n} \geq \hat{\nu}\}, \end{equation} where the threshold $\hat{\nu}$ is chosen so that the false alarm constraint be satisfied with equality and the CUSUM statistic $\{\hat{y}_{n}\}$ admits the following recursion: \begin{equation} \hat{y}_{n}:= (\hat{y}_{n-1})^{+} + \sum_{k=1}^{K} \sum_{j=1}^{b} \left[ \ind{z_{n}^k=j} \log \frac{\mathsf{P}_{0}(z_{n}^k=j)}{\mathsf{P}_{\infty}(z_{n}^k=j)} \right] , ~ \hat{y}_{0}:=0, \label{eq:cus1} \end{equation} Note that we have to multiply by $r$ in \eqref{eq:cus2} in order to return to physical time units, since the samples are acquired with a rate $1/r$. We call this detection scheme Q-CUSUM, where Q stands for the ``quantization'' employed by this method. This detection rule has been studied in \cite{crow}, \cite{mei}, \cite{tart} in the case that the sensors take i.i.d. observations and each sensor communicates with the fusion center at \textit{every} observation time ($r=1$). It is easy to see that as $\gamma \rightarrow \infty$ $$ \frac{\mathcal{J}[\hat{\mathcal{S}}]}{\mathcal{J}[\mathcal{S}]} \rightarrow \frac{r \bar{I}_{0}}{\hat{I}_{0}}, \quad \text{where} \; \hat{I}_{0}:= \frac{1}{K} \sum_{i=1}^{K} \sum_{j=1}^{b} \mathsf{P}_{0}(z_{n}^k=j)\log \frac{ \mathsf{P}_{0}(z_{n}^k=j)}{\mathsf{P}_{\infty}(z_{n}^k=j)} , $$ and $\bar{I}_{0}$ is the average KL-number defined in (\ref{klaver}), which implies that the asymptotic performance of $\hat{\mathcal{S}}$ is optimized by selecting thresholds $\{\Gamma_{j}^k\}$ in order to maximize $\hat{I}_{0}$. However, for any choice of thresholds, $\hat{\mathcal{S}}$ is not (even first-order) asymptotically optimal, since $r \bar{I}_{0}> \hat{I}_{0}$ (see, e.g., \cite{tsi2}). \subsubsection{Fusion of local CUSUM rules} Suppose now that each sensor $k$ communicates at the following times \begin{equation} \label{meitau} \tau_{n}^k= \inf\{ t \geq \tau_{n-1}^k: y^k_{t} \geq c^{k} \} , \end{equation} where $y^k_{t}:= u^k_{t} - \min_{0 \leq s \leq t} u^k_{s}$ is the local CUSUM statistic and $c^{k}$ is a fixed, positive threshold. In this way, the sensors communicate with the fusion center only to announce they have detected the change. This requires only \textit{one-bit} transmissions, which means that even if the network supports the transmission of multi-bit messages, this flexibility is not going to be useful. There are many reasonable fusion center policies that can be based on (\ref{meitau}). For example, the fusion center may raise an alarm the first time any sensor communicates, i.e., at $\min_{k} \tau_{1}^k$ (min-CUSUM). This is clearly a one-shot scheme, i.e., it requires transmission of at most one bit from each sensor, and as one would expect it is asymptotically suboptimal (see, e.g., \cite{tart} for the case of i.i.d. observations and \cite{moustdec} for the case of Brownian motions). An alternative possibility is to raise an alarm the first time that all sensors communicate \textit{simultaneously}, i.e., at $$ \mathcal{M}:= \inf\{ t: y_{t}^k \geq c^{k} , \; \forall \; k=1, \ldots, K\}. $$ This rule was suggested (although in a different form) by Mei \cite{mei}, where it was shown that when each $u^{k}$ is a random walk with a finite second moment, $\mathcal{M}$ is first-order asymptotically optimal (in particular, $\mathcal{J}[\mathcal{M}]- \mathcal{J}[\mathcal{S}]= \mathcal{O}(\sqrt{\log \gamma})$), as long as each $c^{k}$ is proportional to the local KL-number, $I_{0}^k$. Since the constant of proportionality is determined by $\gamma$, this means that for this decentralized scheme, contrary to Q-CUSUM, it is not possible to control how often each sensor communicates with the fusion center. However, by construction, the induced communication activity will be intense only after the change has occurred; before the change, a sensor communicates only to report a ``local false alarm'', which is a rare event. Finally, despite its asymptotic optimality, it is known (see, e.g., \cite{mei}, \cite{tart}) that the non-asymptotic performance of $\mathcal{M}$ can be worse than that of Q-CUSUM when the latter requires transmission of one-bit messages ($b=2$) at every observation time ($r=1$), especially when $K$ is large. \section{D-CUSUM} In this section, we define and analyze the decentralized detection structure that we propose. Thus, we suggest that each sensor $k$ communicates with the fusion center at the following sequence of $\{ {\cF}_{t}^k\}$-stopping times \begin{equation} \label{taucd} \tau_{n}^k := \inf \{ t > \tau_{n-1}^k: u^k_{t} - u^k_{\tau_{n-1}^k} \notin (-\underline{\Delta}^k, \bar{\Delta}^k) \} , ~ n \in \mathbb{N}, \end{equation} where $\tau_0^k:=0$ and $\bar{\Delta}^k, \underline{\Delta}^k$ are fixed, positive thresholds. For every $n \in \mathbb{N}$ and $t>0$ we set $$\tau^{k}(t):=\tau^{k}_{m_{t}^{k}}, \quad m_{t}^k:= \max\{n \in \mathbb{N}: \tau_{n}^k \leq t\}, \quad \ell_{n}^{k}:= u^k_{\tau_{n}^k}- u^k_{\tau_{n-1}^k},$$ i.e., $m_{t}^k$ is the number of messages that have been transmitted by sensor $k$ up to time $t$, $\tau^{k}(t)$ is the most recent communication time for sensor $k$ at time $t$ and $\ell_{n}^{k}$ is the accumulated log-likelihood ratio at sensor $k$ in the time-interval $[\tau_{n-1}^k, \tau^k_{\tau_{n}^k}]$. At time $\tau_n^k$, we suggest that sensor $k$ transmits to the fusion center the following message \begin{equation} \label{zcdmany} z_n^k:=\left\{\begin{array}{cl} j , &\text{if} \quad \bar{\epsilon}^k_{j-1} \leq \ell_{n}^{k} -\bar{\Delta}^k < \bar{\epsilon}^k_{j}\\ -j , &\text{if} \quad -\underline{\epsilon}^k_{j} < \ell_{n}^{k} + \underline{\Delta}^k \leq -\underline{\epsilon}^k_{j-1} \end{array}\right. j=1,\ldots,d, \end{equation} where $\bar{\epsilon}^k_{0}:=\underline{\epsilon}^k_{0}:=0$, $\bar{\epsilon}^k_{d}:=\underline{\epsilon}^k_{d}:= \infty$, $\{\bar{\epsilon}^k_{j}, \underline{\epsilon}^k_{j}\}_{1 \leq j \leq d-1}$ are fixed, positive threshold and $d$ a positive integer. We will also use the following notation $$\bar{\Delta}^k_{j}:=\bar{\Delta}^k + \bar{\epsilon}^k_{j-1} , \quad \underline{\Delta}^k_{j}:=\underline{\Delta}^k + \underline{\epsilon}^k_{j-1}, \quad j=1,\ldots,d,$$ which allows us to rewrite (\ref{zcdmany}) as follows $ z_n^k=\left\{\begin{array}{cl} j , &\text{if} \quad \bar{\Delta}^k_{j} \leq \ell_{n}^{k} < \bar{\Delta}^k_{j+1}\\ -j , &\text{if} \quad -\underline{\Delta}^k_{j+1} < \ell_{n}^{k} \leq -\underline{\Delta}^k_{j} \end{array}\right., \quad j=1,\ldots,d. $$ When $d=1$, $z_{n}^{k}$ is a one-bit message of the form \begin{equation} \label{zcd} z_n^k:=\left\{\begin{array}{cl}1 , &\text{if}~ \ell^k_{n} \geq \bar{\Delta}^k\\ -1 , &\text{if}~ \ell^k_{n} \leq -\underline{\Delta}^k \end{array}\right. \end{equation} that simply informs the fusion center whether $\ell^k_{n} \geq \bar{\Delta}^k$ or $\ell^k_{n}\leq -\underline{\Delta}^k$. When $d \geq 2$, $z_{n}^{k}$ requires the transmission of $\lceil \log_{2}(2d) \rceil=1+ \lceil \log_{2} d\rceil$ bits and the fusion center also obtains information regarding the size of the overshoot. The stopping times (\ref{taucd}) and the messages (\ref{zcdmany}) determine the flow of information (\ref{flow}) at the fusion center. Assuming that the fusion center uses this information and approximates each local log-likelihood ratio $\{u_{t}^{k}\}$ by some statistic $\{\tilde{u}_{t}^{k}\}$, we suggest the following detection rule \begin{equation} \label{dcusum} \dcus:= \inf \{ t \geq 0: \tilde{y}_{t} \geq \tilde{\nu} \}, ~\text{where}~ \tilde{y}_{t}:=\tilde{u}_{t} - \inf_{0 \leq s \leq t} \tilde{u}_{s},\; \tilde{u}_{t}:= \sum_{k=1}^{K} \tilde{u}_{t}^k \end{equation} and threshold $\tilde{\nu}$ is defined so that $\mathsf{E}_{\infty}[-u_{\dcus}]=\gamma$. The appropriate selection for $\tilde{u}_{t}^{k}$, as well as the design and analysis of the resulting detection rule, is different in discrete and continuous time and, for this reason, we will treat these two setups separately. We will see, however, that the proposed detection structure, that we will call D-CUSUM, can be designed in order to have strong asymptotic optimality properties in both cases. \subsection{Continuous-time setup}\label{sec:D-CUSUM-cont} Suppose that each $\{u^{k}_{t}\}$ is a continuous-time process with continuous paths so that condition (\ref{full}) is satisfied, in which case we have the following closed-form expressions for $\mathcal{J}[S]$ and $\gamma$ in terms of threshold $\nu$ (see, e.g., \cite{moustito},\cite{chro}): \begin{align} \label{fapito} \begin{split} \gamma &= \mathsf{E}_{\infty}[-u _{\cus}]= \mathsf{E}_{\infty}[\langle u \rangle_{\cus}] = e^{\nu}-\nu-1 , \\ \mathcal{J}[\cus] &= \mathsf{E}_{0}[u _{\cus}]= \mathsf{E}_{0}[\langle u \rangle_{\cus}]= e^{-\nu}+\nu-1. \end{split} \end{align} Then, each $\ell_{n}^{k}$ is exactly equal to either $\bar{\Delta}^k$ or $-\underline{\Delta}^k$ and, consequently, at $\tau_{n}^k$ sensor $k$ can transmit to the fusion center the \textit{exact} value of $\ell_{n}^{k}$ by simply communicating a \textit{one-bit} message of the form (\ref{zcd}). As a result, the fusion center is able to recover the value of $u^k$ at any time $\tau_{n}^k$, since $u_{\tau_{n}^k}^k = \ell_{1}^{k}+ \ldots+\ell_{n}^{k}$, and a natural approximation for $u_t^k$ at some arbitrary time $t$ is the corresponding most recently reproduced value, i.e., \begin{equation} \label{freecd} \tilde{u}_{t}^k := u^{k}_{\tau^{k}(t)} =\sum_{n=1}^{m_{t}^{k}} \ell_{n}^{k}. \end{equation} The proposed scheme has a number of practical advantages. First of all, the fusion statistic $\{\tilde{y}_{t}\}$ is piecewise-constant and its value needs to be updated only at communication times, according to the following convenient formula: $$ \tilde{y}_{\tau_{n}^{k}}=(\tilde{y}_{\tau_{n}^{k}\text{-}})^{+} + \bar{\Delta}^k\ind{z_n^k=1}-\underline{\Delta}^k\ind{z_n^k=-1}. $$ Compare this with the centralized, continuous-time CUSUM statistic, $\{y_{t}\}$, which does not in general admit such a recursion and whose calculation at the fusion center requires high-frequency transmission of ``infinite-bit'' messages from the sensors. Moreover, it is possible to control the communication rate of sensor $k$ by selecting appropriately $\bar{\Delta}^k$ and $\underline{\Delta}^k$. Since $\mathsf{E}_{i}[\tau_{n}^k -\tau_{n-1}^k]$, $i=0, \infty$ in general depend on $n$, these thresholds can be selected in order to attain target values for $\mathsf{E}_{0}[\ell^k_{n}]$ and $\mathsf{E}_{\infty}[-\ell^k_{n}]$, which do not depend on $n$ and are given by $\mathsf{E}_{0}[\ell^k_{n}]=s(\underline{\Delta}^k , \bar{\Delta}^k)$ and $\mathsf{E}_{\infty}[-\ell^k_{n}] =s(\bar{\Delta}^k,\underline{\Delta}^k)$, where $$s(x,y) := \frac{-x(e^{y}-1)+ye^{y}(e^{x}-1)}{e^{x+y}-1}.$$ In this way, the specification of $\bar{\Delta}^k$ and $\underline{\Delta}^k$ simply requires the solution of a (non-linear) system of two equations. From the previous discussion it should be clear that D-CUSUM is much more preferable than the corresponding centralized CUSUM from a practical point of view. It turns out that it also has excellent performance characteristics, making any additional benefit of the optimal centralized CUSUM test negligible relative to its implementation cost. This becomes clear with the following theorem, which provides a non-asymptotic upper bound on the performance loss of the proposed detection structure. \begin{theorem} \label{prop1} For any $\gamma$ and $\{\bar{\Delta}^k,\underline{\Delta}^k\}_{1 \leq k \leq K}$ we have \begin{equation} \label{order2cd} \mathcal{J}[\dcus] - \mathcal{J}[\cus] \leq 4 \, K \, \Delta_{\max} , \quad \text{where} \quad \Delta_{\max}:= \max_{1 \leq k \leq K} \max \{\bar{\Delta}^k, \underline{\Delta}^k\}. \end{equation} \end{theorem} \begin{proof} The proof is presented in Appendix\,\ref{app:A}. \end{proof} The bound provided in \eqref{order2cd} implies that for any fixed thresholds $\{\bar{\Delta}^k, \underline{\Delta}^k\}$ and any number of sensors $K$, $\mathcal{J}[\dcus] - \mathcal{J}[\cus]=\mathcal{O}(1)$ as $\gamma \rightarrow \infty$, i.e., $\dcus$ is \textit{second}-order asymptotically optimal. In the case of a large sensor-network ($K \rightarrow \infty$), this property is preserved only if we have an asymptotically high rate of communication, specifically if $\Delta_{\max} \rightarrow 0$ so that $K \Delta_{\max} =\mathcal{O}(1)$. However, since we want to avoid intense transmission activity, it is more interesting to see that $\dcus$ remains \textit{first}-order asymptotically optimal when $K \rightarrow \infty$ and $\Delta_{\max} \rightarrow \infty$ so that $K \Delta_{\max}=o(\log \gamma)$. Indeed, from (\ref{fapito}) and (\ref{order2cd}) we have $$ \frac{\mathcal{J}[\dcus]}{\mathcal{J}[\cus]} = 1+ \frac{\mathcal{J}[\dcus] - \mathcal{J}[\cus]}{\mathcal{J}[\cus]} \leq 1+ \frac{4 K \Delta_{\max}}{e^{-\nu}+\nu-1} $$ and our claim now also follows from (\ref{fapito}), which implies that $\nu=\log \gamma+o(1)$. \subsection{Discrete-time setup} Suppose now that each $\{u^{k}_{t}\}$ is a random walk, i.e., the increments $\{u^k_{t}-u_{t-1}^{k}\}_{t \in \mathbb{N}}$ are i.i.d. This implies that each $(\tau_{n}^{k}-\tau_{n-1}^{k}, z_{n}^{k},\ell_{n}^{k})_{n\in \mathbb{N}}$ is a sequence of independent triplets with the same distribution as $(\tau_{1}^{k}, z_{1}^{k},\ell_{1}^{k})$. As a result, thresholds $\bar{\Delta}^k$ and $\underline{\Delta}^k$ can now be selected in order to attain target values for $\mathsf{E}_{i}[\tau_{1}^{k}]$, $i=0,\infty$. However, the main difference with the continuous-time setup is that now each $\ell_{n}^{k}$ is no longer restricted to the binary set $\{\bar{\Delta}^k, -\underline{\Delta}^k\}$. Thus, it now makes sense to have larger than binary alphabets ($d>1$), in which case we also need to select thresholds $\{\bar{\epsilon}^k_{j}, \underline{\epsilon}^k_{j}\}_{1 \leq j \leq d-1}$ (recall that $\bar{\epsilon}^k_{0}=\underline{\epsilon}^k_{0}:=0$, $\bar{\epsilon}^k_{d}=\underline{\epsilon}^k_{d}:= \infty$). We suggest the following specification \begin{align} &\mathsf{P}_0(\ell^k_{1}-\bar{\Delta}^k \geq \bar{\epsilon}^k_j \, | \, \ell_{1}^{k} \geq \bar{\Delta}^k)=1-\frac{j}{d} =\mathsf{P}_\infty(\ell^k_{1}+\underline{\Delta}^k \leq -\underline{\epsilon}^k_j \, | \,\ell_{1}^{k} \leq -\underline{\Delta}^k), \label{eq:levels} \end{align} which guarantees that the overshoot $\ell_{1}^{k} - \bar{\Delta}^k$ (resp. $-(\ell_{1}^{k} + \underline{\Delta}^k)$ is equally likely to lie in each interval $[\bar{\epsilon}^k_{j-1},\bar{\epsilon}^k_{j})$ (resp. $(-\underline{\epsilon}^k_{j}, -\underline{\epsilon}^k_{j-1}]$ given that $\ell_{1}^{k} \geq \bar{\Delta}^k$ (resp. $\ell_{1}^{k} \leq -\underline{\Delta}^k$), i.e., \begin{align*} &\mathsf{P}_0(\ell^k_{1}-\bar{\Delta}^k \in [\bar{\epsilon}^k_{j-1}, \bar{\epsilon}^k_j) \, | \, \ell_{1}^{k} \geq \bar{\Delta}^k)=\frac{1}{d} =\mathsf{P}_\infty(\ell^k_{1}+\underline{\Delta}^k \in (-\underline{\epsilon}^k_{j}, -\underline{\epsilon}^k_{j-1}] \, | \, \ell_{1}^{k} \leq -\underline{\Delta}^k), \end{align*} or, equivalently, $\mathsf{P}_0( z_{1}^{k}=j \, | \, z_{1}^{k}>0 )=1/d =\mathsf{P}_\infty( z_{1}^{k}=-j \, | \, z_{1}^{k}<0 )$, for every $1 \leq j \leq d$. Clearly, all these thresholds can be easily computed off-line, as their computation requires the simulation of the pair $(\tau_{1}^{k}, \ell_{1}^{k})$ under both $\mathsf{P}_{0}$ and $\mathsf{P}_{\infty}$. Moreover, in what follows, we assume that $u_{1}^{k}$ is unbounded and absolutely continuous with a positive density. Then, $\bar{\epsilon}^k_{d-1}, \underline{\epsilon}^k_{d-1} \rightarrow \infty$ as $d \rightarrow \infty$, whereas \begin{equation} \label{del} \epsilon^{k}:= \max_{1 \leq j \leq d-1} \, \{ \bar{\epsilon}^k_{j}-\bar{\epsilon}^k_{j-1} \, , \, \underline{\epsilon}^k_{j}-\underline{\epsilon}^k_{j-1} \} \rightarrow 0 \quad \text{as} \quad d \rightarrow \infty. \end{equation} In order to establish a \textit{second-order} asymptotic optimality property for $\tilde{\mathcal{S}}$, as in the continuous-time setup, we need a lower bound for the optimal centralized performance $\mathcal{J}[\mathcal{S}]$ up to a constant term as $\gamma \rightarrow \infty$. Moreover, in order to obtain the inflicted performance loss as $K \rightarrow \infty$, we need to characterize the growth of this constant term as $K \rightarrow \infty$. This is done in the following lemma, under a second moment condition on each $u_{1}^{k}$. \begin{lemma}\label{lem:6} If $\mathsf{E}_{0}[(u_{1}^{k})^{2}]<\infty$ for every $1 \leq k \leq K$, then for any $\gamma$ we have \begin{equation} \mathcal{J}[\mathcal{S}]= \mathsf{E}_0[u_{\mathcal{S}}]\ge\log\gamma- {\mit \Theta}(K). \end{equation} \end{lemma} \begin{proof} It is well known that the worst case for the optimal centralized CUSUM is when the change occurs at $\tau=0$, which implies the equality in the lemma. The proof of the inequality is presented in Appendix \ref{app:B}. \end{proof} If each sensor $k$ transmitted the exact value of each $\ell_{n}^{k}$ at time $\tau_{n}^{k}$, as in the continuous-time setup, then we could approximate $u_{t}^{k}$ by (\ref{freecd}) and we could work in the same way as Theorem \ref{prop1} to show that $\mathcal{J}[\dcus] -\mathcal{J}[\cus] = \mathcal{O}(K \Delta_{\max})$. However, this is not possible in a discrete-time setup, since $\ell_{n}^{k}$ cannot be fully recovered at the fusion center when sensor $k$ transmits only a small number of bits at time $\tau_{n}^{k}$. Our main goal in the remainder of the paper is to show that it is actually possible to design D-CUSUM in discrete time so that it is second-order asymptotically optimal even if each sensor transmits a small number of bits (such as 2 or 3) in every communication. In order to do this, we approximate $u_{t}^{k}$ by \begin{equation} \label{tilder} \tilde{u}_{t}^k := \sum_{n=1}^{m_{t}^{k}} \tilde{\ell}_{n}^{k}, \end{equation} where $\tilde{\ell}_{n}^{k}$ is the log-likelihood ratio of $z_n^k$, i.e., \begin{align} \tilde{\ell}_{n}^{k} &:= \sum_{j=1}^{d} \Bigl[ \bar{\Lambda}^k_{j} \, \ind{z_{n}^k=j} - \underline{\Lambda}^k_{j} \, \ind{z_{n}^k=-j} \Bigr], \quad \label{ells} \\ \bar{\Lambda}^k_{j} &:= \log \frac{ \mathsf{P}_{0}(z_1^k=j)}{\mathsf{P}_{\infty}(z_1^k=j)}, ~ -\underline{\Lambda}^k_{j} := \log \frac{\mathsf{P}_{0}(z_1^k=-j)}{\mathsf{P}_{\infty}(z_1^k=-j)}. \label{lambdas} \end{align} The log-likelihood ratios $\{\bar{\Lambda}^k_{j},\underline{\Lambda}^k_{j}\}$ do not admit closed-form expressions, however they can be easily computed via simulation. This is not an easy task if one uses their definition in (\ref{lambdas}), which requires simulation of rare events, especially when $\bar{\Delta}^k, \underline{\Delta}^k$ are large. However, we can overcome this problem using the following lemma. \begin{lemma} \label{lem:0} For every $1\leq j \leq d$, $\bar{\Lambda}^k_{j}=\bar{\Delta}^k_{j}+\overline{R}^k_{j}$ and $\underline{\Lambda}^k_{j}=\underline{\Delta}^k_{j}+\underline{R}^k_{j}$, where \begin{align} \label{import} \begin{split} \overline{R}^k_{j} &:= - \log \mathsf{E}_{0}[ e^{-(\ell^k_{1}- \bar{\Delta}^k_{j})} \, | \, z_{1}^{k}=j ] >0 , \\ \underline{R}^k_{j} &:= - \log \mathsf{E}_{\infty}[ e^{\ell^k_{1}+ \underline{\Delta}^k_{j}} \, | \, z_{1}^{k}=-j]>0. \end{split} \end{align} Moreover, for every $1\leq j \leq d-1$, $\overline{R}^k_{j}, \underline{R}^k_{j} \leq \epsilon^{k}$ and if, additionally, $\mathsf{E}_{i}[(u_{1}^{k})^{2}]<\infty$, $i=0, \infty$, then \begin{align} \label{import2} \begin{split} \overline{R}^k_{d} &\leq \mathsf{E}_{0}[ \ell^k_{1}- \bar{\Delta}^k_{d} \, | \, z_{1}^{k}=d ] \leq \Theta(1) \; d \; \mathsf{E}_{0}[(u_{1}^{k})^{2} \ind{u_{1}^{k} \geq \bar{\epsilon}^k_{d-1}}], \\ \underline{R}^k_{d} &\leq \mathsf{E}_{\infty}[ -(\ell^k_{1} + \underline{\Delta}^k_{d}) \, | \, z_{1}^{k}=-d ] \leq \Theta(1) \; d \; \mathsf{E}_{\infty}[(u_{1}^{k})^{2} \ind{-u_{1}^{k} \geq \underline{\epsilon}^k_{d-1}}] , \end{split} \end{align} where $\Theta(1)$ is a term that does not depend on $d$ and is bounded from above and below as $\bar{\Delta}^k, \underline{\Delta}^k \rightarrow \infty$. \end{lemma} \begin{proof} The proof can be found in Appendix\,\ref{app:D}. \end{proof} Lemma \ref{lem:0} shows that, similarly to the thresholds $\{\bar{\Delta}^k_{j},\underline{\Delta}^k_{j}\}$ and $\{\bar{\epsilon}^k_{j},\underline{\epsilon}^k_{j}\}$, the log-likelihood ratios $\{\bar{\Lambda}^k_{j},\underline{\Lambda}^k_{j}\}$ can be computed off-line and efficiently if we simulate $(\tau_{1}^{k}, \ell_{1}^{k})$ under $\mathsf{P}_{0}$ and $\mathsf{P}_{\infty}$. Moreover, Lemma \ref{lem:0} shows that defining $\tilde{\ell}_{n}^{k}$ as the log-likelihood ratio of $z_{n}^{k}$ accounts for the unobserved overshoots at the fusion center. Specifically, when the fusion center receives message $z_{n}^{k}=j$ for some $j=1, \ldots, d$, it understands that $\ell_{n}^{k} \in [\bar{\Delta}^k_{j}, \bar{\Delta}^k_{j+1})$ and it approximates $\ell_{n}^{k}$ by $\bar{\Delta}^k_{j}+ \overline{R}^k_{j}$; in other words, the fusion center approximates the random overshoot $\ell_{n}^{k}-\bar{\Delta}^k_{j}$ that it does not observe by the constant $\overline{R}^k_{j}$, which is clearly an $\mathcal{O}(1)$ term as $\bar{\Delta}^k, \underline{\Delta}^k \rightarrow \infty$. The following lemma is important for quantifying the additional detection delay due to using $\tilde{\ell}_{n}^{k}$ instead of the actual value of $\ell_{n}^{k}$ in (\ref{tilder}). \begin{lemma} \label{lem:1} If $\mathsf{E}_{i}[(u_{1}^{k})^{2}]<\infty$, $i=0, \infty$, then $\mathsf{E}_{0}[\ell^k_1- \tilde{\ell}^k_1 ] \leq 2\theta^{k}$, where \begin{equation} \label{thetak} \theta^{k}:= \epsilon^{k} + \Theta(1) \, \mathsf{E}_{0}[(u_{1}^{k})^{2} \ind{u_{1}^{k} \geq \bar{\epsilon}^k_{d-1}}]+ \Theta(1) \, \mathsf{E}_{\infty}[(u_{1}^{k})^{2} \ind{-u_{1}^{k} \geq \underline{\epsilon}^k_{d-1}}] \end{equation} and $\Theta(1)$ is a term that does not depend on $d$ and is bounded from above and below as $\bar{\Delta}^k, \underline{\Delta}^k \rightarrow \infty$. Moreover, $\theta^{k} \rightarrow 0$ as $d \rightarrow \infty$. \end{lemma} \begin{proof} The proof of this lemma can be found in Appendix\,\ref{app:D}. \end{proof} Note that an alternative approach would have been to define $\tilde{\ell}_{n}^{k}$ as in (\ref{ells}), but with $\bar{\Lambda}^k_{j}$ and $\underline{\Lambda}^k_{j}$ replaced by $\bar{\Delta}^k_{j}$ and $\underline{\Delta}^k_{j}$, respectively. In this way, the overshoots are simply ignored by the fusion center. However, the main reason for defining $\tilde{\ell}_{n}^{k}$ as the log-likelihood ratio of $z_{n}^{k}$ is that it allows us to prove the following lemma, which connects threshold $\tilde{\nu}$ with the false alarm period $\gamma$ and plays a crucial role in establishing the (second-order) asymptotic optimality of the resulting detection rule. \begin{lemma} \label{lem3} For any $\gamma>0$ we have $\tilde{\nu} \leq \log \gamma - \log(\bar{I}_\infty)$, thus, $\tilde{\nu} =\log \gamma +\Theta(1)$ as $\gamma \rightarrow \infty$. \end{lemma} \begin{proof} The proof is presented in Appendix \ref{app:C}. \end{proof} It is possible to prove Lemma \ref{lem3} and, consequently, to establish the asymptotic optimality of $\tilde{S}$ if $\tilde{\ell}_{n}^{k}$ is defined as the log-likelihood ratio of the pair $(\tau_{n}^k-\tau_{n-1}^{k}, z_{n}^{k})$, and not only of $z_{n}^{k}$. Unfortunately, the distribution of $\tau_{1}^k$ is typically intractable, thus, the resulting rule could not be implemented in practice. We are now ready to state the discrete-time analogue of Theorem \ref{prop1}. For simplicity, we assume that communication rates, before and after the change, are of the same order of magnitude for all sensors, i.e., there is a quantity $\Delta$ so that $\bar{\Delta}^k,\underline{\Delta}^k=\Theta(\Delta)$ as $\Delta,\bar{\Delta}^k,\underline{\Delta}^k \rightarrow \infty$ for all $1\leq k \leq K$. Moreover, we set $\theta:=\max_{1 \leq k \leq K} \theta^{k}$. \begin{theorem} \label{th:2} If $\mathsf{E}_{0}[(u_{1}^{k})^{2}]<\infty$ for every $1 \leq k \leq K$, then \begin{equation} \label{order1cd} \mathcal{J}[\dcus] -\mathcal{J}[\cus] \leq \frac{\theta}{\Theta(\Delta)} \, \log\gamma + K \, \Theta(\Delta). \end{equation} \end{theorem} \begin{proof} For the optimum CUSUM $\mathcal{S}$, it is well known that $\mathcal{J}[\mathcal{S}]=\mathsf{E}_0[u_{\mathcal{S}}]$. In order to see that this is also the case for D-CUSUM, i.e., $\mathcal{J}[\tilde{\mathcal{S}}]=\mathsf{E}_0[u_{\tilde{\mathcal{S}}}]$, from the nonnegativity of the KL-divergence it is clear that it suffices to show that $\tilde{S}\ind{\tilde{S} \geq \tau}= \inf\{t \geq\tau: \tilde{y}_{t} \geq \tilde{\nu}\}$ is \textit{pathwise} decreasing with respect to $\tilde{y}_\tau$, or equivalently that the process $\{\tilde{y}_t, t>\tau\}$ is \textit{pathwise} increasing with respect to $\tilde{y}_\tau$. Indeed, if we denote by ($\tau_{n})$ the sequence of times at which there is a communication from at least one sensor, then $$ \tilde{y}_{\tau_n}=(\tilde{y}_{\tau_n-})^++\omega_{\tau_n} $$ where $\omega_{\tau_n}$ is information coming from the sensors that communicate at time $\tau_n$ and is clearly independent from the past. This implies that $\tilde{y}_t$ will be increasing in $(\tilde{y}_\tau)^+$ for any $t \geq \tau$ and our claim follows because the smallest value of the latter quantity is 0. Based on the above, we can write \begin{equation} \mathcal{J}[\tilde{\mathcal{S}}]-\mathcal{J}[\mathcal{S}]=\mathsf{E}_0[u_{\tilde{\mathcal{S}}}]-\mathsf{E}_0[u_{\mathcal{S}}]=\mathsf{E}_0[u_{\tilde{\mathcal{S}}}-\tilde{u}_{\tilde{\mathcal{S}}}] +\mathsf{E}_0[\tilde{u}_{\tilde{\mathcal{S}}}]-\mathsf{E}_0[u_{\mathcal{S}}]. \label{eq:3-part} \end{equation} From Lemma\,\ref{lem:8} we have that $\mathsf{E}_0[\tilde{u}_{\tilde{\mathcal{S}}}] \leq \log \gamma +K \Theta(\Delta)$ and $$\mathsf{E}_0[u_{\tilde{\mathcal{S}}}-\tilde{u}_{\tilde{\mathcal{S}}}] \leq K \Theta (\Delta)+ \theta \, \frac{\log \gamma}{\Theta(\Delta)}.$$ Applying these inequalities and Lemma\,\ref{lem:6} to \eqref{eq:3-part}, we obtain the desired result. Lemma \,\ref{lem:8}, as well as some additional auxiliary results, are stated and proved in Appendix \ref{app:E}. \end{proof} The main consequence of Theorem \ref{th:2} is that D-CUSUM is second-order asymptotically optimal, i.e., $\mathcal{J}[\dcus] -\mathcal{J}[\cus]= \mathcal{O}(1)$, when $K=\mathcal{O}(1)$, $\Delta=\mathcal{O}(1)$ and $\theta \rightarrow 0$ so that $\theta \log\gamma=\mathcal{O}(1)$ as $\gamma \rightarrow \infty$. We have seen in Lemma \ref{lem:1} that $\theta \rightarrow 0$ as $d \rightarrow \infty$. If, in particular, $\theta= \mathcal{O}(1/d^{\alpha})$, where $\alpha$ is some positive constant, then the above analysis implies that $d$ may go to infinity with a rate as low as $\mathcal{O}((\log \gamma)^{1/\alpha})$ and, as a result, the required number of bits per transmission, $1+ \lceil \log_{2} d \rceil$, can be of an order as low as $\mathcal{O}(\frac{1}{\alpha} \log \log \gamma)$. This means that second-order asymptotic optimality is achieved in practice with a very low number of bits per transmission, a conclusion that will also be supported by some simulation experiments in the end of this section. As in continuous time, second-order asymptotic optimality is not preserved with an asymptotically low-rate of communication ($\Delta \rightarrow \infty$). However, from Theorem\,\ref{th:2} and Lemma\,\ref{lem:6} we have \begin{equation} \label{asy2} \frac{\mathcal{J}[\dcus]}{\mathcal{J}[\cus]} =1+ \frac{\mathcal{J}[\dcus] -\mathcal{J}[\cus]}{\mathcal{J}[\cus]} \leq 1+ \frac{ \frac{\theta}{\Theta(\Delta)} + \frac{K \Theta(\Delta)}{\log\gamma}}{1- \frac{\Theta(K)}{\log \gamma}}, \end{equation} which implies that D-CUSUM is first-order asymptotically optimal, i.e., $\mathcal{J}[\dcus]/\mathcal{J}[\cus] \rightarrow 1$, when $\Delta \rightarrow \infty$ so that $K \Delta=o(\log \gamma)$. In this context, the performance of D-CUSUM is optimized when $\Delta, \theta, K$ are selected so that the two terms in the upper bound of \eqref{order1cd} are of the same order magnitude. This happens when $\Delta=\Theta(\sqrt{\theta \log\gamma /K})$, in which case $\mathcal{J}[\dcus] -\mathcal{J}[\cus]= \mathcal{O}(\sqrt{K \, \theta \, \log\gamma})$. We should emphasize that in the case of a binary alphabet ($d=1$), where $\theta$ is bounded away from 0 (i.e., $\theta=\Theta(1)$), first-order asymptotic optimality cannot be achieved with a fixed rate of communication, i.e., when $\Delta=\mathcal{O}(1)$ as $\gamma \rightarrow \infty$. This may seem counterintuitive at first, however it is quite reasonable since a high rate of communication leads to fast accumulation of quantization error. Nevertheless, this source of error can be suppressed if we have a sufficiently large alphabet size that allows us to quantize the overshoots. This explains why first-order asymptotic optimality can be achieved even with $\Delta=\mathcal{O}(1)$ when $\theta \rightarrow 0$. We conclude that, either with a high or a low communication rate, the performance of D-CUSUM is improved with a larger than binary alphabet $(d>1)$, but in practice a small value of $d$ should be sufficient. In order to elaborate more on this point, let us note that the statistical behavior of the overshoots depends on the parameter $\Delta$, which controls the average period of communication in the sensors. However, this dependence is only minor since the distribution of the overshoots converges to some limiting distribution as $\Delta$ becomes large. In other words, quantizing the overshoots is like quantizing a random variable with (almost) fixed statistics. Consequently, the mean square quantization error, or any other similar quality measure, will be (almost) independent from $\Delta$ for fixed number of bits. On the contrary, for the classical quantization scheme \eqref{zdd}, employed by Q-CUSUM, quantization is applied on the value of each $u^k_{nr}-u^k_{(n-1)r}$, where $r$ denotes the fixed corresponding period. It is very simple to realize that for fixed number of bits, if we increase the period $r$, the mean square quantization error will \textit{increase}, since the difference $u^k_{nr}-u^k_{(n-1)r}$ will involve a larger sum of i.i.d.~random variables. This becomes particularly obvious when these random variables are bounded, in which case the support of the sum increases linearly with $r$ and we are asked, with the same number of bits, to quantize a larger range of values. This suggests that if we want to communicate with the fusion center at a smaller rate and preserve the same number of bits, this will inflict larger quantization errors and therefore additional performance degradation for Q-CUSUM. As we mentioned above, this is not the case with the quantization scheme we adopt for D-CUSUM, since increasing $\Delta$ (to reduce the communication rate) leaves the mean square quantization error almost intact. Let us now illustrate these conclusions with a simulation study. Specifically, suppose that each sensor $k$ takes independent, normally distributed observations with variance $1$ and mean that changes from $0$ to $\mu$, i.e., $\xi_{t}^k \sim \mathcal{N}(0,1)$ when $t \leq \tau$ and $\xi_{t}^k \sim \mathcal{N}(\mu,1)$ when $t > \tau$. Then, for every $t \in \mathbb{N}$ we have $u_{t}^k-u_{t-1}^{k} = \mu \, \xi_{t}^k- \mu^{2}/2$. We assume that $\bar{\Delta}^k=\underline{\Delta}^k=\Delta^k$ and for every $j=1, \ldots, d-1$ we set $\bar{\epsilon}^k_{j}=\underline{\epsilon}^k_{j}=\epsilon^k_{j}$ and, consequently, we have $\bar{\Lambda}^k_{j}=\underline{\Lambda}^k_{j}=\Lambda^k_{j}$. Moreover, we assume that each $\Delta^{k}$ is chosen so that $\mathsf{E}_{0}[\tau_{1}^{k}]=r$. In Table\,\ref{tab:1} we present the values of these parameters when the number of transmitted bits per message is $d=1$ or $d=2$, the communication period is $r=3$ or $r=6$ and $\mu=1$. \begin{table}[!h] \centering \caption{Thresholds and Log-Likelihood Ratios} \begin{tabular}{cc} \begin{tabular}{|c||c|c|c|c|} \hline & $\Delta^k_{1}$ & $\Lambda^k_{1}$ \\ \hline \hline $r=3$, $\mu=1$ & 1.287 & 1.87 \\ \hline $r=6$, $\mu=1$ & 2.54 & 3.12 \\ \hline \end{tabular} & \begin{tabular}{|c||c|c|c|c|}\hline & $\Delta^k_{1}$ & $\Delta^k_{2}$ & $\Lambda^k_{1}$ & $\Lambda^k_{2}$ \\ \hline \hline $r=3$, $\mu=1$ & 1.287 & 1.87 & 1.54 & 2.94 \\ \hline $r=6$, $\mu=1$ & 2.54 & 3.12 & 2.80 & 3.62 \\ \hline \end{tabular}\\ &\\ (a) $\;d=1$ & (b) $\;d=2$ \end{tabular} \label{tab:1} \end{table} \begin{comment} \begin{table}[!h] \centering \begin{tabular}{c|c|c|c|c|c|c} & $\Delta^k_{1}$ & $\Delta^k_{2}$ & $\Delta^k_{3}$ & $\Lambda^k_{1}$ & $\Lambda^k_{2}$ & $\Lambda^k_{3}$ \\ \hline $\delta=3$, $\mu=0.5$ & 0.55 & 0.72 & 0.95 & 0.63 & 0.98 & 1.18 \\ \hline $\delta=6$, $\mu=0.5$ & 0.99 & 1.15 & 1.37 &1.06 & 1.40 & 1.61 \\ \hline $\delta=3$, $\mu=1$ & 1.287 & 1.647 & 2.15 & 1.45 & 2.23 & 3.46 \\ \hline $\delta=6$, $\mu=1$ & 2.54 & 2.90 & 3.41 & 2.71 & 3.47 & 3.85 \\ \hline $\delta=3$, $\mu=2$ & 4.29 & 5.17 & 6.37 & 4.69 & 6.55 & 7.21 \\ \hline $\delta=6$, $\mu=2$ & 10.33 & 11.25 & 12.36 &10.68 &12.55 & 13.22 \\ \hline \end{tabular} \caption{The case $d=3$} \label{tab:1} \end{table} \end{comment} \begin{comment} More specifically, when $\mu=1$, we need to choose $\Delta=1.29$ in order to have $r=3$ and $\Delta=2.55$ in order to have $r=6$. Moreover, when $d=1$, it will be $\Lambda_{1}=1.87$ when $r=3$ and $\Lambda_{1}=3.12$ when $r=6$. On the other hand, when $d=2$, it is $\epsilon_{2}=0.57$ for both $r=3$ and $r=6$, whereas it is $\Lambda_{1}=1.54$, $\Lambda_{2}=2.35$ for $r=3$ and $\Lambda_{1}=2.80$, $\Lambda_{2}=3.61$ for $r=6$. When $\mu=0.5$, we must choose $\Delta=0.56$ for $r=3$ and $\Delta=0.99$ for $r=6$. When $d=1$, it will be $\Lambda_{1}=0.85$ when $r=3$ and $\Lambda_{1}=1.28$ when $r=6$. On the other hand, when $d=2$, it is $\epsilon_{2}=0.25$ for both $r=3$ and $r=6$, whereas it is $\Lambda_{1}=0.67$ , $\Lambda_{2}=1.06$ for $r=3$ and $\Lambda_{1}=1.10$, $\Lambda_{2}=1.49$ when $r=6$. \end{comment} Our goal is to compare D-CUSUM $\tilde{\mathcal{S}}$ with Q-CUSUM $\hat{\mathcal{S}}$, which was defined in (\ref{eq:cus2}), when both rules use the same resources, i.e., the same number of bits per communication and the same (average) rate of communication. Note that such a fair comparison is not possible with decentralized rules that do not explicitly control their transmission rate. Of course, the ultimate benchmark is the centralized CUSUM test, which requires transmission of the observation of each sensor at every time $t$. \begin{figure}[!ht] \centering \includegraphics{fig1.pdf} \caption{Case of $K=5$ sensors with communication period $r=3$.} \label{fig:1} \vskip0.5cm \centering \includegraphics{fig2.pdf} \caption{Case of $K=5$ sensors with communication period $r=6$.} \label{fig:2} \end{figure} Fig.\,\ref{fig:1} and Fig.\,\ref{fig:2} depict the main results of our simulations. First of all, we observe that in all cases the operating characteristic curve of D-CUSUM $\dcus$ is essentially parallel to that of the optimal centralized CUSUM, $\mathcal{S}$. This is exactly the \textit{second-order} asymptotic optimality that we established theoretically. On the contrary, the operating characteristic curve of Q-CUSUM $\hat{\mathcal{S}}$ diverges as $\gamma$ increases, as expected, since this not an asymptotically optimal scheme (even of first order). Of course, when an ``infinite-bit'' message is transmitted at each communication time, Q-CUSUM corresponds to the centralized CUSUM with period $r$ and its operating characteristic curve is parallel to the optimal one. However, what is really interesting is that D-CUSUM with one-bit or two-bit transmissions is either very close or even outperforms this \textit{infinite-bit} Q-CUSUM. Finally, we should also note that when the average communication period is small ($r=3$), there is a considerable improvement in D-CUSUM when using two, instead of one, bits per transmission (see Fig.\,\ref{fig:1}). On the other hand, when the average communication period is large ($r=6$), we do not observe similar performance gains for D-CUSUM by having the sensors transmit additional bits to the fusion center (see Fig.\,\ref{fig:2}). \section{Conclusions} The main contribution of this paper is a novel decentralized sequential detection rule, that we called D-CUSUM, according to which each sensor communicates with the fusion center at two-sided exit times of its local log-likelihood ratio and the fusion center uses in parallel a CUSUM-like rule in order to detect the change. We showed that the performance loss of D-CUSUM with respect to the optimal centralized CUSUM is bounded as the rate of false alarms goes to 0 (\textit{second order} asymptotic optimality). Moreover, we showed that its first-order asymptotic optimality is preserved even with an asymptotically low communication rate and large number of sensors. We illustrated these properties with simulation experiments, which also showed that D-CUSUM performs significantly better than a CUSUM-based, decentralized detection rule that requires communication at deterministic times. We assumed throughout the paper that observations from different sensors are independent, an assumption which is not needed for the optimality of the centralized CUSUM test, but is universal in the decentralized literature. This assumption is necessary both for the design and the analysis of D-CUSUM in discrete time, however it is possible to remove it in continuous time, at least when the sensors observe \textit{correlated} Brownian motions. Indeed, going over the proof of Theorem \ref{prop1} in Appendix\,\ref{app:A}, we realize that this assumption is needed only to the extent that it guarantees a decomposition of the form $u_{t}= \sum_{k=1}^{K} u_{t}^k$, where $\{u^k_{t}\}$ is an $ {\cF}_{t}^k$-adapted process with continuous paths. That is, we did not use explicitly the fact that $\{u^k_{t}\}$ is the local log-likelihood ratio at sensor $k$. This implies that Theorem \ref{prop1} will remain valid even for sensors with correlated dynamics, as long as such a decomposition is possible. This is indeed the case when the sensors observe correlated Brownian motions before and after the change, i.e., for every $1 \leq k \leq K$ it is $$ \xi^k_{t}= \sum_{j=1}^{K} \sigma_{kj} W^{j}_{t} + \ind{t > \tau} \, \mu^k t, ~ t \geq 0,$$ where $(W^{1}, \ldots,W^{K})$ is a standard $K$-dimensional Wiener process, $\mu=[\mu^{1}, \ldots, \mu^{K}]'$ a $K$-dimensional real vector and $\sigma:=[\sigma_{ij}]$ a square matrix of dimension $K$ so that the diffusion coefficient matrix ${\mit\Sigma}= \sigma \sigma'$ is invertible. Then, we can write $u_{t}= \sum_{k=1}^{K} [ b^k \, \xi_{t}^k - 0.5 \, \mu^k \, b^k \, t]$, where $b=[b^1,\ldots,b^K]'= {\mit\Sigma}^{-1} \mu$, and Theorem\,\ref{prop1} remains valid as long as we define $u^{k}_{t}$ in (\ref{taucd}) not as the local log-likelihood ratio $\mu^k \, \xi_{t}^k - 0.5 \, (\mu^k)^{2} \, t$, but as $b^k \xi_{t}^k - 0.5 \, \mu^k b^k t$. However, it remains an open problem to establish asymptotically optimal, decentralized detection rules for more general continuous-time models, and of course in the i.i.d. setup, when the sensor observations are correlated.
\section{Introduction} Random metric theory is based on the idea of randomizing the classical space theory of functional analysis. This idea may date back to K. Menger, B. Schweizer and A. Sklar's idea of the theory of probabilistic metric spaces, cf. \cite{SS}. Random normed modules (briefly, $RN$ modules), random inner product modules (briefly, $RIP$ modules) and random locally convex modules (briefly, $RLC$ modules) have been the basic framework in random metric theory, and the theory of random conjugate spaces has been a powerful tool for the deep development of the basic framework. Random metric theory may now also be aptly called random functional analysis since it is being developed as functional analysis over the basic random framework. Motivated by the original notions of random metric spaces and random normed spaces introduced in \cite{SS}, all the basic notions such as $RN$, $RIP$ and $RLC$ modules together with their random conjugate spaces were naturally presented by Guo in the course of the development of random functional analysis, cf. \cite{TXG-Master,TXG-PHD,TXG-Extension,TXG-Module,TXG-Radon,TXG-basic,TXG-Sur,TXG-JFA}. Recently, we are pleased to learn that several other authors also ever independently presented the notions of $RN$ modules and random conjugate spaces. For example, as a tool for the study of ultrapowers of Lebesgue-Bochner function spaces, R.Haydon, M.Levy and Y.Raynaud introduced real $RN$ modules (called randomly normed $L^{0}$--modules in \cite{HLR}) and their random conjugate spaces, cf. \cite{HLR}. Motivated by financial applications, D. Filipovi\'{c}, M. Kupper and N. Vogelpoth also presented real $RN$ modules (called $L^{0}$--normed modules in \cite{FKV}), and in particular the notions of locally $L^{0}$--convex modules and the locally $L^{0}$--convex topology for random locally convex modules, cf. \cite{FKV}. Before 2009, all the theory of $RN$ modules and $RLC$ modules together with their random conjugate spaces was developed under the $(\varepsilon,\lambda)$--topology. The $(\varepsilon,\lambda)$--topology is inherited from B.Schweizer and A. Sklar's work in 1960, where they first introduced such a kind of topology for more abstract probabilistic metric spaces. The $(\varepsilon,\lambda)$--topology is an abstract generalization of the topology of convergence in measure on the linear space of random variables, and hence the essence of the $(\varepsilon,\lambda)$--topology is locally nonconvex in general. Therefore, from a viewpoint of topological modules, all the pre-2009 work on random metric theory can be regarded a locally nonconvex generalization of classical results established for normed spaces and locally convex spaces, in the process the theory of random conjugate spaces has played an essential role and the order structure and module structure peculiar to $RN$ and $RLC$ modules not only can be fully available for the theory of random conjugate spaces but also lead to the difficulties in the study of complicated stratification structure. In 2009, in \cite{FKV} D.Filipovi$\acute{c}$, M.Kupper and N.Vogelpoth presented the notion of a locally $L^{0}$--convex module, in company with which the notion of a locally $L^{0}$--convex topology for an $RN$ module and $RLC$ module was also introduced. This means that the $L^{0}$--norm on an $RN$ module or the family of $L^{0}$--seminorms on an $RLC$ module can also induce another kind of topology, namely the locally $L^{0}$--convex topology. Subsequently, the principal connections between some basic results derived from the two kinds of topologies for an $RN$ module or $RLC$ module were given in \cite{TXG-JFA}, based on which, lots of new and basic researches recently have been done in \cite{TXG-GS,TXG-YJY,TXG-XLZ,TXG-XZ,TXG-SEZ,MZW,SEZ-TXG}. The recent study further exposes the respective advantages and disadvantages of the two kinds of topologies. Although the locally $L^{0}$--convex topology is very similar to the classical locally convex topology, the study of the locally $L^{0}$--convex topology often requires the involved $L^{0}$--modules or their subsets to have the countable concatenation property, namely the stratification structure of a locally $L^{0}$--convex module has a remarkable effect on its topological structure, which makes the theory of the locally $L^{0}$--convex topology considerably different from that of the classical locally convex topology. Therefore, we have been hoping that combining the respective advantages of the two kinds of topologies produces a perfect random convex analysis over random locally convex modules for better financial applications, since D. Filipovi\'{c}, et.al's paper \cite{FKV} only consider the corresponding problems under the locally $L^{0}$--convex topology. The central purpose of this paper is to achieve such a goal. Classical convex analysis (e.g, see \cite{ET}) is the analytic foundation for convex risk measures, cf. \cite{ADEH,Delbaen,FS,Follmer-S,Fritt-R}. However, it is no longer universally applicable to $L^{0}$--convex (or conditional convex) conditional risk measures (in particular, those defined on the model spaces of unbounded financial positions). Just to overcome the obstacle, D. Filipovi\'{c}, et.al presented a good idea of randomizing the initial data, which leads to an attempt to establish random convex analysis over locally $L^{0}$--convex modules in order to provide a new approach to conditional risk, called the module approach, cf, \cite{FKV,FKV-appro}. Although the notions of locally $L^{0}$--convex modules and in particular the locally $L^{0}$--convex topology are very important, we recently find out that many foundational problems induced by their paper \cite{FKV} are worth further studying and perfecting, which also makes our work in this paper urgent. First, the premise for most of the principal results in \cite{FKV} is that the locally $L^{0}$--convex topology for every locally $L^{0}$--convex module can be induced by a family of $L^{0}$--seminorms. Although they ever gave a proof of the premise (namely, Theorem 2.4 of \cite{FKV}), there was a hole in the proof of \cite{FKV}, even according to some remarks in this paper we think that this premise may be not valid in general. In this paper we choose random locally convex modules as the framework on which random convex analysis is based so that we can avoid the theoretically difficult point involved in \cite{FKV} since the definition of a random locally convex module assumes the existence of a family of $L^{0}$--seminorms in advance and $L^{0}$--seminorms are often easily constructed in both the theoretic study and financial applications. Second, Lemma 2.28 of \cite{FKV} played a crucial role in the proofs of the main results-Proposition 3.4, Theorems 2.8 and 3.8 of \cite{FKV}. However, in this paper we provide a counterexample to both Lemma 2.28 and Theorem 2.8 of \cite{FKV}, thus the two results and the related results should be modified, which is done in Subsections 3.1 and 3.2 of this paper, and in particular the refined version of the modified results can also be obtained since the precise relation between random conjugate spaces of a random locally convex module under the two kinds of topologies has been found in this paper, in particular the refined results will also be needed in the sequel of this paper. Third, although D. Filipovi\'{c}, M. Kupper and N. vogepoth had presented the notion of $L^{0}$--barreled modules and established the continuity and subdifferentiability theorems for $L^{0}$--convex functions defined on $L^{0}$--barreled modules, namely Proposition 3.5 and Theorem 3.7 of \cite{FKV}, the two results can not be applied to the study of conditional risk measures since it remains open up to now whether the model space $L^{p}_{\mathcal{F}}(\mathcal{E})$ playing a crucial role in the module approach to conditional risk is an $L^{0}$--barreled module. In this paper, we overcome the difficulty by presenting the notion of an $L^{0}$--pre-barreled module. The notion of an $L^{0}$--pre-barreled module is weaker than that of an $L^{0}$--barreled module and meets the needs of financial applications. To prove this, we establish random duality theory of a random duality pair under random locally convex modules endowed with the locally $L^{0}$--convex topology so that we can give a characterization for random locally convex modules to be $L^{0}$--pre-barreled modules, in particular $L^{p}_{\mathcal{F}}(\mathcal{E})$ is $L^{0}$--pre-barreled when it is endowed with the locally $L^{0}$--convex topology. Further, we also establish the new continuity and subdifferentiability theorems based on the notion of an $L^{0}$--pre-barreled module. All these results are given in Subsections 3.3 and 3.4. Concerning random convex analysis over random locally convex modules under the $(\varepsilon,\lambda)$--topology, although it is impossible to establish the corresponding continuity and subdifferentiability theorems under the $(\varepsilon,\lambda)$--topology since the $(\varepsilon,\lambda)$--topology is locally nonconvex in general, we can give a natural Fenchel-Moreau dual representation theorem, which has the same shape as the classical Fenchel-Moreau dual representation theorem. Since the $(\varepsilon,\lambda)$--topology is in harmony with the norm topology, for example, let $(\Omega,\mathcal{E},P)$ be a probability space, $\mathcal{F}$ a $\sigma$--subalgebra of $\mathcal{E}$ and $L^{p}_{\mathcal{F}}(\mathcal{E})$ the $RN$ module generated by the Banach space $L^{p}(\mathcal{E})$ $(1\leq p\leq+\infty)$, then $L^{p}(\mathcal{E})$ is dense in $L^{p}_{\mathcal{F}}(\mathcal{E})$ with respect to the $(\varepsilon,\lambda)$--topology on $L^{p}_{\mathcal{F}}(\mathcal{E})$ (clearly, this is not true with respect to the locally $L^{0}$--convex topology!). The simple fact motivates us to futher study the precise relations among the three kinds of conditional risk measures. The first kind was introduced independently by K.Detlefsen and G.Scandolo in \cite{DS} and J.Bion-Nadal in \cite{Nadal} as a monotone and cash-invariant function from $L^{\infty}(\mathcal{E})$ to $L^{\infty}(\mathcal{F})$ (briefly, an $L^{\infty}$--conditional risk measure). The second and third kinds were introduced by D. Filipovi\'{c}, M. Kupper and N. Vogelpoth in \cite{FKV-appro} as monotone and cash-invariant functions from $L^{p}(\mathcal{E})$ to $L^{r}(\mathcal{F})$ $(1\leq r\leq p<+\infty)$ and from $L^{p}_{\mathcal{F}}(\mathcal{E})$ to $\bar{L}^{0}(\mathcal{F})$ $(1\leq p\leq+\infty)$, respectively, (briefly, $L^{p}$-- and $L^{p}_{\mathcal{F}}(\mathcal{E})$--conditional risk measures, respectively). We show that an $L^{\infty}$--conditional risk measure can be uniquely extended to an $L^{\infty}_{\mathcal{F}}(\mathcal{E})$--conditional risk measure and the conditional convex dual representation theorem established in \cite{Nadal,DS} for the former can be regarded as a special case of that established in this paper for the latter. We further show that an $L^{0}$--convex $L^{p}$--conditional risk measure can be uniquely extended to an $L^{0}$--convex $L^{p}_{\mathcal{F}}(\mathcal{E})$--conditional risk measure $(1\leq p<+\infty)$ and the conditional convex dual representation theorem established in \cite{FKV-appro} for the former can also be regarded a special case of that established in \cite{FKV-appro} for the latter. The second extension theorem is not very easy, whose proof is constructive, since an $L^{0}$--convex $L^{p}$--conditional risk measure is not necessarily uniformly continuous with respect to the relative topology when $L^{p}(\mathcal{E})$ is regarded as a subspace of $L^{p}_{\mathcal{F}}(\mathcal{E})$ which is endowed with the $(\varepsilon,\lambda)$--topology, in particular, in the process we find that combining the countable concatenation hull of a set and the local property of conditional risk measures is a very useful analytic skill that may considerably simplify and improve the study of $L^{0}$--convex conditional risk measures. This shows that the two vector space approaches to conditional risk can be incorporated into the module approach. Thus this paper has provided a complete random convex analysis, and hence also a solid analytic foundation for the module approach to conditional risk. Most of the main ideas and results of this paper were first announced in Guo's survey paper \cite{TXG-Recent} without the detailed proofs or at most with a sketch of proofs of a few of illustrative results, this paper includes many new results as well as the full proofs of the results announced in \cite{TXG-Recent}. Besides. the great distinction between this paper and \cite{TXG-Recent} is that almost all the results in Section 3 is stated under the framework of random locally convex modules endowed with the locally $L^{0}$--convex topology rather than the framework of locally $L^{0}$--convex modules for some reasons mentioned above. The remainder of this paper is organized as follows. Section 2 provides some necessary preliminaries for sake of the reader's convenience and in particular includes some key new results on the precise relations between the two kinds of closures of an $L^{0}$--convex set and between the two kinds of random conjugate spaces of a random locally convex module under the $(\varepsilon,\lambda)$--topology and locally $L^{0}$--convex topology; Sections 3 and 4 present and prove our main results as stated above in the Introduction of this paper. Throughout this paper, we always use the following notation and terminology: $K:$ the scalar field R of real numbers or C of complex numbers. $(\Omega,\mathcal{F},P):$ a probability space. $L^{0}(\mathcal{F},K)=$ the algebra of equivalence classes of $K$--valued $\mathcal{F}$-- measurable random variables on $(\Omega,\mathcal{F},P)$. $L^{0}(\mathcal{F})=L^{0}(\mathcal{F},R)$. $\bar{L}^{0}(\mathcal{F})=$ the set of equivalence classes of extended real-valued $\mathcal{F}$-- measurable random variables on $(\Omega,\mathcal{F},P)$. As usual, $\bar{L}^{0}(\mathcal{F})$ is partially ordered by $\xi\leq\eta$ iff $\xi^{0}(\omega)\leq\eta^{0}(\omega)$ for $P$--almost all $\omega\in \Omega$ (briefly, a.s.), where $\xi^0$ and $\eta^0$ are arbitrarily chosen representatives of $\xi$ and $\eta$, respectively. Then $(\bar{L}^{0}(\mathcal{F}),\leq)$ is a complete lattice, $\bigvee H$ and $\bigwedge H$ denote the supremum and infimum of a subset $H$, respectively. $(L^{0}(\mathcal{F}),\leq)$ is a conditionally complete lattice. Please refer to \cite{D-Sch} or \cite[p. 3026]{TXG-JFA} for the rich properties of the supremum and infimum of a set in $\bar{L}^{0}(\mathcal{F})$. Let $\xi$ and $\eta$ be in $\bar{L}^{0}(\mathcal{F})$. $\xi<\eta$ is understood as usual, namely $\xi\leq\eta$ and $\xi\neq\eta$. In this paper we also use $``\xi<\eta $ (or $\xi\leq\eta$) on $A"$ for $``\xi^{0}(\omega)<\eta^{0}(\omega)$ (resp., $\xi^{0}(\omega)\leq\eta^{0}(\omega)$) for $P$--almost all $\omega\in A"$, where $A\in\mathcal{F}$, $\xi^0$ and $\eta^0$ are a representative of $\xi$ and $\eta$, respectively. $\bar{L}^{0}_{+}(\mathcal{F})=\{\xi\in \bar{L}^{0}(\mathcal{F})~|~\xi\geq0\}$ $\L^{0}_{+}(\mathcal{F})=\{\xi\in L^{0}(\mathcal{F})~|~\xi\geq0\}$ $\bar{L}^{0}_{++}(\mathcal{F})=\{\xi\in \bar{L}^{0}(\mathcal{F})~|~\xi>0$ on $\Omega\}$ $L^{0}_{++}(\mathcal{F})=\{\xi\in L^{0}(\mathcal{F})~|~\xi>0$ on $\Omega\}$ Besides, $\tilde{I}_{A}$ always denotes the equivalence class of $I_{A}$, where $A\in \mathcal{F}$ and $I_{A}$ is the characteristic function of $A$. When $\tilde{A}$ denotes the equivalence class of $A (\in \mathcal{F})$, namely $\tilde{A}=\{B\in\mathcal{F}~|~P(A\triangle B)=0\}$ (here, $A\triangle B=(A\setminus B)\bigcup(B\setminus A)$), we also use $I_{\tilde{A}}$ for $\tilde{I}_{A}$. Specially, $[\xi<\eta]$ denotes the equivalence class of $\{\omega\in\Omega~|~\xi^0(\omega)<\eta^0(\omega)\}$, where $\xi^0$ and $\eta^0$ are arbitrarily chosen representatives of $\xi$ and $\eta$ in $\bar{L}^0(\mathcal{F})$, respectively, some more notations such as $[\xi=\eta]$ and $[\xi\neq\eta]$ can be similarly understood. \section{Preliminaries} \subsection{Some basic notions} \begin{definition}($See$ \cite{TXG-Master,TXG-PHD,TXG-basic}.) An ordered pair $(E,\|\cdot\|)$ is called a random normed space (briefly, an $RN$ space) over $K$ with base $(\Omega,\mathcal{F},P)$ if $E$ is a linear space over $K$ and $\|\cdot\|$ is a mapping from $E$ to $L^{0}_{+}(\mathcal{F})$ such that the following are satisfied: \noindent ($RN$--1). $\|\alpha x\|=|\alpha| \|x\|$, $\forall \alpha\in K$ and $x\in E$; \noindent ($RN$--2). $\|x\|=0$ implies $x=\theta$ (the null element of $E$); \noindent ($RN$--3). $\|x+y\|\leq\|x\|+\|y\|$, $\forall x,y\in E$. \noindent Here $\|\cdot\|$ is called the random norm on $E$ and $\|x\|$ the random norm of $x\in E$ (If $\|\cdot\|$ only satisfies ($RN$--1) and ($RN$--3) above, it is called a random seminorm on $E$). \noindent Furthermore, if, in addition, $E$ is a left module over the algebra $L^{0}(\mathcal{F},K)$ (briefly, an $L^{0}(\mathcal{F},K)$--module) such that \noindent ($RNM$--1). $\|\xi x\|=|\xi| \|x\|$, $\forall \xi\in L^{0}(\mathcal{F},K)$ and $x\in E$. \noindent Then $(E,\|\cdot\|)$ is called a random normed module (briefly, an $RN$ module) over $K$ with base $(\Omega,\mathcal{F},P)$, the random norm $\|\cdot\|$ with the property ($RNM$--1) is also called an $L^0$--norm on $E$ (a mapping only satisfying ($RN$--3) and ($RNM$--1) above is called an $L^0$--seminorm on E). \end{definition} \begin{remark} According to the original notion of an $RN$ space in \cite{SS}, $\|x\|$ is a nonnegative random variable for all $x\in E$. An $RN$ space in the sense of Definition 2.1 is almost equivalent to (in fact, slightly more general than) the original one in the sense of \cite{SS}. Definition 2.1 is not only very natural from traditional functional analysis but also easily avoids any possible ambiguities between random variables and their equivalence classes, and hence also more convenient for applications to Lebesgue-Bochner function spaces since the latter exactly consists of equivalence classes. $RN$ spaces in the sense of Definition 2.1 was essentially earlier employed in \cite{TXG-Master}. The study of random conjugate spaces (see Definition 2.3 below) of $RN$ spaces and applications of $RN$ spaces to best approximations in Lebesgue-Bochner function spaces lead Guo to the notion of an $RN$ module in \cite{TXG-PHD}. Subsequently, $RN$ modules and their random conjugate spaces were deeply developed by Guo in \cite{TXG-PHD,TXG-Extension,TXG-Module,TXG-Radon,TXG-reflesive,TXG-dual} so that Guo further presented the refined notions of $RN$ and random inner product (briefly, $RIP$) modules and compared the original notions of random metric spaces (briefly, $RM$ spaces) and $RN$ spaces with the currently used notions of $RM$ and $RN$ spaces in \cite{TXG-basic}. At almost the same time as Guo did the work \cite{TXG-Master,TXG-PHD}, $RN$ spaces and $RN$ modules were independently introduced by R. Haydon, M. Levy and Y. Raynaud in \cite{HLR}, where their notion of randomly normed $L^0(\mathcal{F})$--modules is exactly that of $RN$ modules over $R$ with base $(\Omega,\mathcal{F},P)$, in particular, they deeply studied the two classes of $RN$ modules-direct integrals and random Banach lattices (namely, random normed module equivalent of Banach lattices). Motivated by financial applications, D. Filipovi\'{c}, M. Kupper and N. Vogelpoth also independently came to the idea of $RN$ modules, their notion of $L^0$--normed modules amounts to that of $RN$ modules over $R$ with base $(\Omega,\mathcal{F},P)$. The terminologies of ``$L^0$--norms and $L^0$--seminorms" are adopted from \cite{FKV}, previously they were still called random norms and random seminorms in Guo's work and \cite{HLR}. \end{remark} \begin{definition}($See$ \cite{TXG-Master,TXG-PHD,TXG-Module,TXG-basic}.) Let $(E_1,\|\cdot\|)$ and $(E_2,\|\cdot\|)$ be $RN$ spaces over $K$ with base $(\Omega,\mathcal{F},P)$. A linear operator $T$ from $E_1$ to $E_2$ is said to be a.s. bounded if there is $\xi\in L^{0}_{+}(\mathcal{F})$ such that $\|Tx\|_2\leq\xi\|x\|_1, \forall x\in E_1$. Denote by $B(E_1,E_2)$ the linear space of a.s. bounded linear operators from $E_1$ to $E_2$, define $\|\cdot\|:B(E_1,E_2)\rightarrow L^{0}_{+}(\mathcal{F})$ by $\|T\|=\bigwedge\{\xi\in L^{0}_{+}(\mathcal{F})~|~\|Tx\|_2\leq\xi\|x\|_1$ for all $x\in E_1\}$ for all $T\in B(E_1,E_2)$, then it is easy to check that $(B(E_1,E_2),\|\cdot\|)$ is also an $RN$ space over $K$ with base $(\Omega,\mathcal{F},P)$, in particular $(B(E_1,E_2),\|\cdot\|)$ is an $RN$ module if so is $E_2$. Specially, the $RN$ module $(E_1^{\ast},\|\cdot\|)$ with $E_1^{\ast}=B(E_1,L^{0}(\mathcal{F},K))$ is called the random conjugate space of $E_1$. \end{definition} \begin{remark} In Definition 2.3, let $A$ be a subalgebra dense in $L^0(\mathcal{F})$ with respect to the topology of convergence in measure, then it is easy to prove that every a.s. bounded linear operator $T$ between two $RN$ $A$--modules is an $A$--homomorphisms, thus in this case $B(E_1,E_2)$ is the same as in \cite{HLR}. Here, the notion of an $RN$ $A$--module was introduced in \cite{HLR}, which can obtained by replacing $L^0(\mathcal{F})$ with $A$ in the definition of an $RN$ module. Although $RN$ spaces and $RN$ $A$--modules are both more general than $RN$ modules, the history of random metric theory has proved that $RN$ modules are most important. \end{remark} . \begin{definition}($See$ \cite{TXG-PHD,TXG-Module,TXG-Sur}.) An ordered pair $(E,\mathcal{P})$ is called a random locally convex space (briefly, an $RLC$ space) over $K$ with base $(\Omega,\mathcal{F},P)$ if $E$ is a linear space over $K$ and $\mathcal{P}$ a family of mappings from $E$ to $L^0_{+}(\mathcal{F})$ such that the following are satisfied: \noindent ($RLC$--1). Every $\|\cdot\|\in \mathcal{P}$ is a random seminorm on $E$; \noindent ($RLC$--2). $\bigvee\{\|x\|:\|\cdot\|\in\mathcal{P}\}=0$ iff $x=\theta$. \noindent Furthermore, if, in addition, $E$ is an $L^{0}(\mathcal{F},K)$--module and each $\|\cdot\|\in \mathcal{P}$ is an $L^0$--seminorm on $E$, then $(E,\mathcal{P})$ is called a random locally convex module (briefly, an $RLC$ module) over $K$ with base $(\Omega,\mathcal{F},P)$. \end{definition} It is not very difficult to introduce the random conjugate space for an $RN$ space, whereas it is completely another thing to do for an $RLC$ space, at the earlier time Guo ever gave two definitions. \begin{definition}($See$ \cite{TXG-PHD,TXG-Module,TXG-Sur}.) Let $(E,\mathcal{P})$ be an $RLC$ space over $K$ with base $(\Omega,\mathcal{F},P)$. A linear operator $f$ from $E$ to $L^0(\mathcal{F},K)$ (such an operator is also called a random linear functional on $E$) is called an a.s. bounded random linear functional of type I if there are $\xi\in L^0_+(\mathcal{F})$ and some finite subset $\mathcal{Q}$ of $\mathcal{P}$ such that $|f(x)|\leq\xi \|x\|_{\mathcal{Q}}$ for all $x\in E$, where $\|\cdot\|_{\mathcal{Q}}=\bigvee\{\|\cdot\|:\|\cdot\|\in \mathcal{Q}\}$, namely, $\|x\|_{\mathcal{Q}}=\bigvee\{\|x\|:\|\cdot\|\in \mathcal{Q}\}$ for all $x\in E$. Denote by $E^{\ast}_{I}$ the $L^0(\mathcal{F},K)$--module of a.s. bounded random linear functionals on $E$ of type I, called the first kind of random conjugate space of $(E,\mathcal{P})$, cf. \cite{TXG-PHD,TXG-Module}. A random linear functional $f$ on $E$ is called an a.s. bounded random linear functional of type II if there are $\xi\in L^0_+(\mathcal{F})$ and $\|\cdot\|\in \mathcal{P}_{cc}$ such that $|f(x)|\leq\xi\|x\|$ for all $x\in E$, where $\mathcal{P}_{cc}=\{\sum_{n=1}^{\infty}\tilde{I}_{A_n}\|\cdot\|_{\mathcal{Q}_n}~|~\{A_n,n\in N\}$ is a countable partition of $\Omega$ to $\mathcal{F}$ and $\{\mathcal{Q}_n,n\in N\}$ a sequence of finite subsets of $\mathcal{P}\}$. Denote by $E^{\ast}_{II}$ the $L^0(\mathcal{F},K)$--module of a.s. bounded random linear functionals on $E$ of type II, called the second kind of random conjugate space of $(E,\mathcal{P})$, cf. \cite{TXG-Sur}. \end{definition} In the sequel of this paper, given a random locally convex space $(E,\mathcal{P})$, $\mathcal{P}_f$ always denotes the family of finite subsets of $\mathcal{P}$, $\|\cdot\|_{\mathcal{Q}}$ is the same as in Definition 2.6 for each $\mathcal{Q}\in \mathcal{P}_f$ and $\mathcal{P}_{cc}$ also the same as in Definition 2.6. $\mathcal{P}_{cc}$ is called the countable concatenation hull of $\mathcal{P}$. Following are the three important examples in random metric theory. \begin{example} Let $L^0(\mathcal{F},B)$ be the $L^0(\mathcal{F},K)$--module of equivalence classes of $\mathcal{F}$--random variables (or, strongly $\mathcal{F}$--measurable functions) from $(\Omega,\mathcal{F},P)$ to a normed space $(B,\|\cdot\|)$ over $K$. $\|\cdot\|$ induces an $L^0$--norm (still denoted by $\|\cdot\|$) on $L^0(\mathcal{F},B)$ by $\|x\|:=$ the equivalence class of $\|x^0(\cdot)\|$ for all $x\in L^0(\mathcal{F},B)$, where $x^0(\cdot)$ is a representative of $x$. Then $(L^0(\mathcal{F},B),\|\cdot\|)$ is an $RN$ module over $K$ with base $(\Omega,\mathcal{F},P)$. Specially, $L^{0}(\mathcal{F},K)$ is an $RN$ module, the $L^0$--norm $\|\cdot\|$ on $L^{0}(\mathcal{F},K)$ is still denoted by $|\cdot|$. \end{example} $L^0(\mathcal{F},B)$ was deeply studied by Guo in \cite{TXG-PHD,TXG-Module,TXG-Radon,TXG-SBL,TXG-JAT,ZYY-TXG}. When the norm $\|\cdot\|$ on $B$ is not fixed, let $\{\|\cdot\|_\omega,\omega\in\Omega\}$ be an $\mathcal{F}$--measurable family of norms (namely $\|b\|_\omega$ is an $\mathcal{F}$--random variable as a function of $\omega\in\Omega$ for each fixed $b\in B$) and $B_\omega=$ the completion of $(B,\|\cdot\|_\omega)$ for each $\omega\in\Omega$. R. Haydon, et.al, introduced the notion of a generalized strongly $\mathcal{F}$--measurable function in \cite{HLR}, where an element $f$ in the product space $\Pi_{\omega\in\Omega}B_{\omega}$ was called a generalized strongly $\mathcal{F}$--measurable function on $\Omega$ if there is a sequence $\{f_n,n\in N\}$ of $B$--valued $\mathcal{F}$--simple functions on $\Omega$ such that $\|f(\omega)-f_n(\omega)\|_\omega\rightarrow 0 (n\rightarrow\infty)$ for each $\omega\in\Omega$. Denote by $\int^{\oplus}_{\Omega}B_\omega P(d\omega)$ the $L^0(\mathcal{F},K)$--module of equivalence classes of generalized strongly $\mathcal{F}$--measurable functions on $\Omega$, define the $L^0$--norm $\|\cdot\|$ by $\|x\|=$ the equivalence class of the mapping sending each $\omega$ to $\|x^0(\omega)\|_\omega$, where $x^0$ is a representative of $x\in\int^{\oplus}_{\Omega}B_\omega P(d\omega)$, then $(\int^{\oplus}_{\Omega}B_\omega P(d\omega),\|\cdot\|)$ is an $RN$ module over $K$ with base $(\Omega,\mathcal{F},P)$, called the direct integral of $\{B_\omega,\omega\in\Omega\}$, which played a key role in \cite{HLR}. D. Filipovi$\acute{c}$, M. Kupper and N. Vogelpoth constructed important $RN$ modules $L^p_{\mathcal{F}}(\mathcal{E}) (1\leq p\leq+\infty)$ in \cite{FKV}, we will prove that they play the role of universal model spaces for $L^0$--convex conditional risk measures. \begin{example} Let $(\Omega, {\mathcal E}, P)$ be a probability space and ${\mathcal F}$ a $\sigma$--subalgebra of ${\mathcal E}$. Define $|||\cdot|||_p\colon L^0({\mathcal E})\to {\bar L}^0_+({\mathcal F})$ by $$|||x|||_p=\left\{ \begin{array}{ll} E[|x|^p|{\mathcal F}]^{1\over p}, & \hbox{when $1\leq p<\infty$;} \\ \bigwedge\{\xi\in {\bar L}^0_+({\mathcal F})~|~|x|\leq\xi\}, & \hbox{when $p=+\infty$;} \end{array} \right. $$ for all $x\in L^0({\mathcal E})$. Denote $L^p_{\mathcal F}({\mathcal E})=\{x\in L^0({\mathcal E})~|~|||x|||_p\in L^0_+({\mathcal F})\}$, then $(L^p_{\mathcal F}({\mathcal E}), |||\cdot|||_p)$ is an $RN$ module over $R$ with base $(\Omega, {\mathcal F}, P)$ and $L^p_{\mathcal F}({\mathcal E})=L^0({\mathcal F})\cdot L^p({\mathcal E})=\{~\xi x~|~\xi\in L^0({\mathcal F})~ \hbox{and}~ x\in L^p({\mathcal E})\}$. \end{example} To put some important classes of stochastic processes into the framework of $RN$ modules, Guo constructed a more general $RN$ module $L^p_{\mathcal F}(S)$ in \cite{TXG-JFA} for each $p\in [1, +\infty]$, one can imagine that $S$ is an $RN$ module generated by a class of stochastic processes, $L^p_{\mathcal F}(S)$ can be constructed as follows. \begin{example} Let $(S, \|\cdot\|)$ be an $RN$ module over $K$ with base $(\Omega, {\mathcal E}, P)$ and ${\mathcal F}$ a $\sigma$--subalgebra. Define $|||\cdot|||_p\colon S \to {\bar L}^0_+({\mathcal F})$ by $$|||x|||_p=\left\{ \begin{array}{ll} E[\|x\|^p|{\mathcal F}]^{1\over p}, & \hbox{when $1\leq p<\infty$;} \\ \bigwedge\{\xi\in {\bar L}^0_+({\mathcal F})|~\|x\|\leq\xi\}, & \hbox{when $p=+\infty$;} \end{array} \right. $$ for all $x\in S$. Denote $L^p_{\mathcal F}(S)=\{x\in S~|~|||x|||_p\in L^0_+({\mathcal F})\}$, then $(L^p_{\mathcal F}(S), |||\cdot|||_p)$ is an $RN$ module over $K$ with base $(\Omega, {\mathcal F}, P)$. When $S=L^0({\mathcal E})$, $L^p_{\mathcal F}(S)$ is exactly $L^p_{\mathcal F}({\mathcal E})$. \end{example} \begin{definition}($See$ \cite{TXG-PHD,TXG-Module,TXG-Sur,TXG-SLP}.) Let $(E, {\mathcal P})$ be an $RLC$ space over $K$ with base $(\Omega, {\mathcal F}, P)$. For any positive numbers $\varepsilon$ and $\lambda$ with $0<\lambda<1$ and $\mathcal{Q}\in {\mathcal P}_f$, let $N_{\theta}(\mathcal{Q}, \varepsilon, \lambda)=\{x\in E~|~P\{\omega\in \Omega~|~\|x\|_\mathcal{Q}(\omega)<\varepsilon\}>1-\lambda\}$, then $\{N_{\theta}(\mathcal{Q}, \varepsilon, \lambda)~|~\mathcal{Q}\in {\mathcal P}_f, \varepsilon >0, 0<\lambda<1\}$ forms a local base at $\theta$ of some Hausdorff linear topology on $E$, called the $(\varepsilon, \lambda)$--topology induced by ${\mathcal P}$. \end{definition} From now on, we always denote by ${\mathcal T}_{\varepsilon, \lambda}$ the $(\varepsilon, \lambda)$--topology for every $RLC$ space if there is no possible confusion. Clearly, the $(\varepsilon, \lambda)$--topology for the special class of $RN$ modules $L^0({\mathcal F}, B)$ is exactly the ordinary topology of convergence in measure, and $(L^0({\mathcal F}, K), {\mathcal T}_{\varepsilon, \lambda})$ is a topological algebra over $K$. It is also easy to check that $(E, {\mathcal T}_{\varepsilon, \lambda})$ is a topological module over $(L^0({\mathcal F}, K), {\mathcal T}_{\varepsilon, \lambda})$ when $(E, {\mathcal P})$ is an $RLC$ module over $K$ with base $(\Omega, {\mathcal F}, P)$, namely the module multiplication operation is jointly continuous. For an $RLC$ module $(E, {\mathcal P})$ over $K$ with base $(\Omega, {\mathcal F}, P)$, we always denote by $(E, {\mathcal P})^\ast_{\varepsilon, \lambda}$ ( or, briefly, $E^\ast_{\varepsilon, \lambda}$, whenever there is no confusion ) the $L^0({\mathcal F}, K)$--module of continuous module homomorphisms from $(E, {\mathcal T}_{\varepsilon, \lambda})$ to $(L^0({\mathcal F}, K), {\mathcal T}_{\varepsilon, \lambda})$, called the random conjugate space of $(E, {\mathcal P})$ under the $(\varepsilon, \lambda)$--topology. \begin{proposition}($See$ \cite{TXG-PHD,TXG-Module}.) Let $(E_1, \|\cdot\|_1)$ and $(E_2, \|\cdot\|_2)$ be two $RN$ modules over $K$ with base $(\Omega, {\mathcal F}, P)$ and $T$ a linear operator from $E_1$ to $E_2$. Then $T\in B(E_1, E_2)$ iff $T$ is a continuous module homomorphism from $(E_1, {\mathcal T}_{\varepsilon, \lambda})$ to $(E_2, {\mathcal T}_{\varepsilon, \lambda})$, in which case $\|T\|=\bigvee\{\|Tx\|_2~|~x\in E_1~\hbox{and}~ \|x\|_1\leq 1\}$. \end{proposition} Proposition 2.11 is very useful, Guo use it to prove that $(B(E_1, E_2), \|\cdot\|)$ is always ${\mathcal T}_{\varepsilon, \lambda}$--complete for any two $RN$ spaces $E_1$ and $E_2$ such that $E_2$ is ${\mathcal T}_{\varepsilon, \lambda}$--complete, in particular $E^\ast$ is ${\mathcal T}_{\varepsilon, \lambda}$--complete for every $RN$ space $E$, cf. \cite{TXG-PHD,TXG-Module}. It is also clear from Proposition 2.11 that $E^\ast=E^\ast_{\varepsilon, \lambda}$ for every $RN$ module $E$, cf. \cite{TXG-PHD,TXG-Extension}. Proposition 2.11 can be extended to a more general case when $E_1$ and $E_2$ are $RLC$ modules, cf. \cite{TXG-Sur,TXG-LHZ}. But, this paper only needs a special case of the general result in \cite{TXG-Sur,TXG-LHZ}, namely Proposition 2.12 below. \begin{proposition}($See$ \cite{TXG-Sur,TXG-LHZ}.) Let $(E, {\mathcal P})$ be an $RLC$ module $(E, {\mathcal P})$ over $K$ with base $(\Omega, {\mathcal F}, P)$ and $f$ a random linear functional on $E$. Then $f\in E^\ast_{II}$ iff $f\in E^\ast_{\varepsilon, \lambda}$, namely $E^\ast_{II}=E^\ast_{\varepsilon, \lambda}$. \end{proposition} For any $\varepsilon \in L^0_{++}({\mathcal F})$, let $U(\varepsilon)=\{\xi\in L^0({\mathcal F}, K)~|~|\xi|\leq \varepsilon\}$. A subset $G$ of $L^0({\mathcal F}, K)$ is ${\mathcal T}_c$--open if for each fixed $x\in G$ there is some $\varepsilon \in L^0_{++}({\mathcal F})$ such that $x+U(\varepsilon)\subset G$. Denote by ${\mathcal T}_c$ the family of ${\mathcal T}_c$--open subsets of $L^0({\mathcal F}, K)$, then ${\mathcal T}_c$ is a Hausdorff topology on $L^0({\mathcal F}, K)$ such that $(L^0({\mathcal F}, K), {\mathcal T}_c)$ is a topological ring, namely the addition and multiplication operations are jointly continuous. D. Filipovi\'{c}, M. Kupper and N. Vogelpoth first observed this kind of topology and further pointed out that ${\mathcal T}_c$ is not necessarily a linear topology since the mapping $\alpha\mapsto \alpha x$ ($x$ is fixed) is no longer continuous in general. These observations led them to the study of a class of topological modules over the topological ring $(L^0({\mathcal F}, K), {\mathcal T}_c)$ in \cite{FKV}, where they only considered the case when $K=R$, in fact the complex case can also similarly introduced as follows. \begin{definition}($See$ \cite{FKV}.) An ordered pair $(E, {\mathcal T})$ is a topological $L^0({\mathcal F}, K)$--module if both $(E, {\mathcal T})$ is a topological space and $E$ is an $L^0({\mathcal F}, K)$--module such that $(E, {\mathcal T})$ is a topological module over the topological ring $(L^0({\mathcal F}, K), {\mathcal T}_c)$, namely the addition and module multiplication operations are jointly continuous. \end{definition} Denote by $(E, {\mathcal T})^\ast_c$ ( briefly, $E^\ast_c$ ) the $L^0({\mathcal F}, K)$--module of continuous module homomorphisms from $(E, {\mathcal T})$ to $(L^0({\mathcal F}, K), {\mathcal T}_c)$, called the random conjugate space of the topological $L^0({\mathcal F}, K)$--module $(E, {\mathcal T})$, which was first introduced in \cite{FKV}. \begin{definition}($See$ \cite{TXG-Sur,TXG-XXC,FKV}.) Let $E$ be an $L^0({\mathcal F}, K)$--module and $A$ and $B$ two subsets of $E$. $A$ is said to be $L^0$--absorbed by $B$ if there is some $\xi \in L^0_{++}({\mathcal F})$ such that $\eta A\subset B$ for all $\eta\in L^0({\mathcal F}, K)$ with $|\eta|\leq \xi$. $B$ is $L^0$--absorbent if $B$ $L^0$--absorbs every element in $E$. $B$ is $L^0$--convex if $\xi x+(1-\xi)y\in B$ for all $x,\,y\in B$ and $\xi\in L^0_+({\mathcal F})$ with $0\leq \xi\leq 1$. $B$ is $L^0$--balanced if $\eta B\subset B$ for all $\eta\in L^0({\mathcal F}, K)$ with $|\eta|\leq 1$. \end{definition} \begin{remark} Clearly, when $B$ is $L^0$--balanced, $A$ is $L^0$--absorbed by $B$ iff there exists some $\xi \in L^0_{++}({\mathcal F})$ such that $A\subset \xi B$. Since $L^0({\mathcal F}, K)$ is an algebra over $K$, an $L^0({\mathcal F}, K)$--module is also a linear space over $K$, then it is clear that $B$ is balanced (resp., convex) if $B$ is $L^0$--balanced (resp., $L^0$--convex). But `` being $L^0$--absorbent '' and `` being absorbent '' may not imply each other. \end{remark} \begin{definition}($See$ \cite{FKV}.) A topological $L^0({\mathcal F}, K)$--module $(E, {\mathcal T})$ is called a locally $L^0$--convex $L^0({\mathcal F}, K)$--module ( briefly, a locally $L^0$--convex module when $K=R$ ), in which case ${\mathcal T}$ is called a locally $L^0$--convex topology on $E$, if ${\mathcal T}$ has a local base ${\mathcal B}$ at $\theta$ ( the null element in $E$ ) such that each member in ${\mathcal B}$ is $L^0$--balanced, $L^0$--absorbent and $L^0$--convex. \end{definition} \begin{definition}($See$ \cite{FKV}.) Let ${\mathcal P}$ be a family of $L^0$--seminorms on an $L^0({\mathcal F}, K)$--module $E$. For any $\varepsilon \in L^0_{++}({\mathcal F})$ and any $ Q\in {\mathcal P}_f$ (namely $Q$ is a finite subset of ${\mathcal P}$), let $N_{\theta}(Q, \varepsilon)=\{x\in E~|~\|x\|_Q\leq \varepsilon\}$, then $\{~N_{\theta}(Q, \varepsilon)~|~Q\in {\mathcal P}_f, ~\varepsilon \in L^0_{++}({\mathcal F})\}$ forms a local base at $\theta$ of some locally $L^0$--convex topology, called the locally $L^0$--convex topology induced by ${\mathcal P}$. \end{definition} \begin{corollary} Let $(E,{\mathcal P})$ be an $RLC$ module over $K$ with base $(\Omega, {\mathcal F}, P)$ and ${\mathcal T}_c$ the locally $L^0$--convex topology induced by ${\mathcal P}$. Then $(E, {\mathcal T}_c)$ is a Hausdorff locally $L^0$--convex $L^0({\mathcal F}, K)$--module. \end{corollary} From now on, we always denote by $\mathcal{T}_c$ the locally $L^0$--convex topology induced by $\mathcal{P}$ for every $RLC$ module $(E,\mathcal{P})$ if there is no risk of confusion. Let $(E,{\mathcal P})^\ast_c=(E,{\mathcal T}_c)^\ast_c$ (briefly, $E^\ast_c$, if there is no risk of confusion), called the random conjugate space of a random locally convex module $(E,{\mathcal P})$ under the locally $L^0$--convex topology ${\mathcal T}_c$ induced by ${\mathcal P}$. \begin{proposition}($See$ \cite{TXG-JFA}.) Let $(E,{\mathcal P})$ be a random locally convex module over $K$ with base $(\Omega, {\mathcal F}, P)$ and $f\colon E\to L^0({\mathcal F}, K)$ a random linear functional. Then $f\in E^\ast_I$ iff $f\in E^\ast_c$, namely $E^\ast_I=E^\ast_c$. \end{proposition} \begin{remark} In \cite{TXG-JFA}, it is proved that $ E^\ast_c\subset E^\ast_I$ ( see \cite[p.3032]{TXG-JFA} ). Conversely, if $f\in E^\ast_I$, namely $f$ is a random linear functional and there are some $\xi\in L^0_+({\mathcal F})$ and $Q\in {\mathcal P}_f$ such that $|f(x)|\leq \xi \|x\|_Q$ for all $x\in E$. Lemma 2.12 of \cite{TXG-JFA} shows that $f$ must be $L^0({\mathcal F}, K)$--linear. It is also clear that $f$ is continuous from $(E,{\mathcal T}_c)$ to $(L^0({\mathcal F}, K), {\mathcal T}_c)$, and hence $f\in E^\ast_c$. Thus Proposition 2.19 has been proved in \cite{TXG-JFA}. \end{remark} A family ${\mathcal P}$ of random seminorms on a linear space $E$ is said to be countably concatenated (or to have the countable concatenation property) if ${\mathcal P}_{cc}={\mathcal P}$, this definition appears stronger than that given in \cite{FKV} for a family of $L^0$--seminorms on an $L^0({\mathcal F}, K)$--module since ${\mathcal P}$ must be invariant under the operation of finitely many suprema once ${\mathcal P}_{cc}={\mathcal P}$. But ${\mathcal P}$ and $\{~\|\cdot\|_Q:Q\in {\mathcal P}_f\}$ always induces the same locally $L^0$--convex topology for any family ${\mathcal P}$ of $L^0$--seminorms on an $L^0({\mathcal F}, K)$--module $E$, thus the definition is essentially equivalent to that introduced in \cite{FKV}. It is also obvious that $E^\ast_I=E^\ast_{II}$ if ${\mathcal P}_{cc}={\mathcal P}$, and hence we have the following: \begin{corollary} ($See$ \cite{TXG-JFA}.) Let $(E,{\mathcal P})$ be an $RLC$ module over $K$ with base $(\Omega, {\mathcal F}, P)$. Then $E^\ast_c=E^\ast_{\varepsilon, \lambda}$ if ${\mathcal P}$ has the countable concatenation property. Specially, $E^\ast=E^\ast_c=E^\ast_{\varepsilon, \lambda}$ for an $RN$ module $(E,\|\cdot\|)$. \end{corollary} For an $RLC$ module $(E, {\mathcal P})$, by definition $E^\ast_I\subset E^\ast_{II}$, so $E^\ast_c\subset E^\ast_{\varepsilon, \lambda}$ by Propositions 2.12 and 2.19, see Subsection 2.3 for the precise relation between $E^\ast_c$ and $E^\ast_{\varepsilon, \lambda}$. In the final part of this subsection, let us return to the basic problem: whether can a locally $L^0$--convex topology on an $L^0({\mathcal F}, K)$--module $E$ be induced by a family of $L^0$--seminorms on $E$? If the answer is yes, then the theory of Hausdorff locally $L^0$--convex modules is equivalent to that of random locally convex modules endowed with the locally $L^0$--convex topology, which will be a perfect counterpart of the classical result that a Hausdorff locally convex topology can be equivalently expressed by a separating family of seminorms. It is well known that classical gauge functionals play a crucial role in the proof of the classical result. Let $U$ be a balanced, absorbent and convex subset of a locally convex space $(E, {\mathcal T})$ and $p_U$ the gauge functional of $U$, then the following relation is easily verified: $$U^o\subset \{x\in E~|~p_U(x)<1\}\subset U\subset \{x\in E~|~p_U(x)\leq 1\}, \eqno(2.1)$$ It is the relation (2.1) that is key in the proof of the above classical result. Random gauge functional was first introduced in \cite{FKV}. Let $U$ be an $L^0$--balanced, $L^0$--absorbent and $L^0$--convex subset of an $L^0({\mathcal F}, K)$--module $E$, define $p_U\colon E\to L^0_+({\mathcal F})$ by $p_U(x)=\bigwedge\{\xi\in L^0_+({\mathcal F})~|~x\in \xi U\}$ for all $x\in E$, called the random gauge functional of $U$. Furthermore, it is also proved in \cite{FKV} that $p_U(x)=\bigwedge\{\xi\in L^0_{++}({\mathcal F})~|~x\in \xi U\}$ for all $x\in E$ and $p_U$ is an $L^0$--seminorm on $E$. If, in addition, let $(E, {\mathcal T})$ be a locally $L^0$--convex $L^0({\mathcal F}, K)$--module, D. Filipovi\'{c}, M. Kupper and N. Vogelpoth already proved in \cite{FKV} the following: \begin{proposition}($See$ \cite{FKV}.) Let $(E, {\mathcal T})$ be a locally $L^0$--convex $L^0({\mathcal F}, K)$--module and $U$ an $L^0$--balanced, $L^0$--absorbent and $L^0$--convex subset of $E$. Then the following statements hold:\\ (i).~~$U\subset \{x\in E~|~p_U(x)\leq 1\}$;\\ (ii).~~$p_U(x)\geq 1 $ on $B$ if ${\tilde I}_Ax\notin {\tilde I}_AK$ for all $A\in {\mathcal F}$ with $P(A)>0$ and $A\subset B$, where $B\in {\mathcal F}$ satisfies $P(B)>0$;\\ (iii).~~$U^o\subset \{x\in E~|~p_U(x)<1~\hbox{on}~\Omega\}$. \end{proposition} The most interesting part in Proposition 2.22 is (ii). In fact, (i) is clear and (iii) can be proved as follows: Given an $x\in U^o$, there is an $L^0$--balanced, $L^0$--absorbent and $L^0$--convex neighborhood $V$ of $\theta$ such that $x+V\in U$. Since there is $\delta \in L^0_{++}({\mathcal F})$ such that $\delta x\in V$, $(1+\delta)x=x+\delta x\in x+V\subset U$, so $x\in {\frac{1}{1+\delta}}U$, then $p_U(x)\leq {\frac{1}{1+\delta}}<1~\hbox{on}~\Omega$. Then, can (ii) of Proposition 2.22 imply that $\{x\in E~|~p_U(x)< 1~\hbox{on}~\Omega\}\subset U$? Or, we can ask: does it hold that $\{x\in E~|~p_U(x)< 1~\hbox{on}~\Omega\}\subset U$? If it can not be guaranteed that $\{x\in E~|~p_U(x)< 1~\hbox{on}~\Omega\}\subset U$, then it is still not clear whether Theorem 2.4 of \cite{FKV} is valid, namely whether a locally $L^0$--convex topology can be induced by a family of $L^0$--seminorms. Proposition 2.24 below shows that it may be not a simple problem whether $\{x\in E~|~p_U(x)< 1~\hbox{on}~\Omega\}$ is contained in $U$, from this we even conjuncture that a locally $L^0$--convex topology may not necessarily be induced by a family of $L^0$--seminorms. Let us first recall the notion of countable concatenation property of a set or an $L^0({\mathcal F}, K)$--module. The introducing of the notion utterly results from the study of the locally $L^0$--convex topology, the reader will see that this notion is ubiquitous in the theory of the locally $L^0$--convex topology. From now on, we always suppose that all the $L^0({\mathcal F}, K)$--modules $E$ involved in this paper have the property that for any $x,~y\in E$, if there is a countable partition $\{A_n,n\in N\}$ of $\Omega$ to ${\mathcal F}$ such that ${\tilde I}_{A_n}x={\tilde I}_{A_n}y$ for each $n\in N$ then $x=y$. Guo already pointed out in \cite{TXG-JFA} that all random locally convex modules possess this property, so the assumption is not too restrictive. \begin{definition}($See$ \cite{TXG-JFA}.) Let $E$ be an $L^0({\mathcal F}, K)$--module. A sequence $\{x_n, n\in N\}$ in $E$ is countably concatenated in $E$ with respect to a countable partition $\{A_n,n\in N\}$ of $\Omega$ to ${\mathcal F}$ if there is $x\in E$ such that ${\tilde I}_{A_n}x={\tilde I}_{A_n}x_n$ for each $n\in N$, in which case we define $\sum^{\infty}_{n=1}{\tilde I}_{A_n}x_n$ as $x$. A subset $G$ of $E$ is said to have the countable concatenation property if each sequence $\{x_n, n\in N\}$ in $G$ is countably concatenated in $E$ with respect to an arbitrary countable partition $\{A_n,n\in N\}$ of $\Omega$ to $\mathcal{F}$ and $\sum^{\infty}_{n=1}{\tilde I}_{A_n}x_n\in G$. \end{definition} \begin{proposition} Let $(E, {\mathcal T})$ be a locally $L^0$--convex $L^0({\mathcal F}, K)$--module and $U$ an $L^0$--balanced, $L^0$--absorbent and $L^0$--convex subset with the countable concatenation property. Then $U^o\subset \{x\in E~|~p_U(x)<1 ~\hbox{on}~\Omega\}\subset U\subset \{x\in E~|~p_U(x)\leq 1\}$, where $U^o$ denotes the ${\mathcal T}$--interior of $U$. \end{proposition} \begin{proof} By Proposition 2.22, we only need to show that $\{x\in E~|~p_U(x)<1 ~\hbox{on}~\Omega\}\subset U$. Let $x_{0}$ be a point in $E$ such that $p_U(x_{0})<1~\hbox{on}~\Omega$. Since $\{\xi\in L^0_{++}({\mathcal F})~|~x_{0}\in \xi U\}$ is downward directed, there is a sequence $\{\xi_n, n\in N\}$ in $L^0_{++}({\mathcal F})$ such that it converges to $p_U(x_{0})$ in a nonincreasing way and $x_{0}\in \xi_n U$ for each $n\in N$. By the Egoroff theorem there are a countable partition $\{A_n,n\in N\}$ of $\Omega$ to ${\mathcal F}$ and a subsequence $\{\xi_{n_k},k\in N\}$ of $\{\xi_n, n\in N\}$ such that the subsequence converges to $p_U(x_{0})$ uniformly on each $A_n$. So, we can suppose that the subsequence is just $\{\xi_n, n\in N\}$ itself and each $\xi_n<1~\hbox{on}~A_n$ since $p_U(x_{0})<1 ~\hbox{on}~\Omega$. Clearly, ${\tilde I}_{A_n}x_{0}\in {\tilde I}_{A_n}\xi_n U$ for each $n\in N$, let $u_n\in U$ be such that ${\tilde I}_{A_n}x_{0}={\tilde I}_{A_n}\xi_n u_n$ for each $n\in N$ and $\xi=\sum^{\infty}_{n=1}{\tilde I}_{A_n}\xi_n$, then there is $u\in U$ such that $u=\sum^{\infty}_{n=1}{\tilde I}_{A_n}u_n$ since $U$ has the countable concatenation property. Since $x_{0}=\sum^{\infty}_{n=1}{\tilde I}_{A_n}x_{0}=\sum^{\infty}_{n=1}{\tilde I}_{A_n}\xi_n u_n=\xi u$, $x_{0}\in \xi U\subset U$ by noticing that $\xi<1~\hbox{on}~\Omega$ and $U$ is an $L^0$--convex set with $\theta\in U$.\hfill $\square$ \end{proof} \begin{remark} We can also give another shorter proof of Proposition 2.24 as follows: If $x_{0}\notin U$, then there is a set $B\in {\mathcal F}$ with $P(B)>0$ such that ${\tilde I}_A x_{0}\notin {\tilde I}_A U$ for all $A\in {\mathcal F}$ such that $A\subset B$ and $P(A)>0$ by Theorem 3.13 of \cite{TXG-JFA} since $U$ has the countable concatenation property. Then $p_U(x_{0})\geq 1~\hbox {on}~B$ by (ii) of Proposition 2.22. \end{remark} The above two proofs both use the countable concatenation property of $U$. Thus we ask: Does Proposition 2.24 hold if $U$ lacks the countable concatenation property? Recently, in \cite{TXG-GS} we proved that $E$ must have the countable concatenation property if a random locally convex module $(E,{\mathcal P})$ has a ${\mathcal T}_c$--open neighborhood $U$ of $\theta$ such that $U$ has the countable concatenation property (it is also known in \cite{TXG-GS} that at this time the ${\mathcal T}_c$--interior $U^o$ of $U$ has the countable concatenation property if $U$ is a subset such that $U^o\neq\emptyset$ and $U$ has the countable concatenation property), then we naturally ask: If $E$ has the countable concatenation property, then may a locally $L^0$--convex topology ${\mathcal T}$ on $E$ have a local base ${\mathcal B}$ at $\theta$ such that each $U\in {\mathcal B}$ is an $L^0$--balanced, $L^0$--absorbent and $L^0$--convex subset with the countable concatenation property? We conjuncture that the answers to the above two problems are both negative according to our our experience from our recent work. Thus we are forced to frequently work with the framework of random locally convex modules $(E, {\mathcal P})$ since only the framework can fully develop the power of both the family ${\mathcal P}$ of $L^0$--seminorms and the locally $L^0$--convex topology ${\mathcal T}_c$ induced by ${\mathcal P}$, which has been enough both for the theoretic development and for financial applications as far as the work of this paper is concerned. \subsection{On some precise connections between the $(\varepsilon, \lambda)$--topology and the locally $L^0$--convex topology for a random locally convex module} In \cite{TXG-JFA}, Guo already proved the following two results, which will be used in this paper. \begin{proposition}($See$ \cite{TXG-JFA}.) Let $(E, {\mathcal P})$ be an $RLC$ module. Then $E$ is ${\mathcal T}_{\varepsilon, \lambda}$--complete iff both $E$ has the countable concatenation property and $E$ is ${\mathcal T}_c$--complete. \end{proposition} \begin{proposition}($See$ \cite{TXG-JFA}.) Let $(E, {\mathcal P})$ be an $RLC$ module and $G$ a subset of $E$ such that $G$ has the countable concatenation property. Then ${\bar G}_{\varepsilon, \lambda}={\bar G}_c$, where ${\bar G}_{\varepsilon, \lambda}$ and ${\bar G}_c$ denotes the ${\mathcal T}_{\varepsilon, \lambda}$-- and ${\mathcal T}_c$-- closures of $G$, respectively. \end{proposition} Theorem 2.28 below will play a key role in the proof of the random bipolar theorem in Subsection 3.3.1 of this paper. \begin{theorem} Let $(E, {\mathcal P})$ be an $RLC$ module over $K$ with base $(\Omega, {\mathcal F}, P)$ such that $E$ has the countable concatenation property. Then the following are true:\\ \noindent (1). ${\bar G}_c={\bar G}_{\varepsilon, \lambda}$ has the countable concatenation property if so does $G$;\\ \noindent (2). If $G$ is $L^0$--convex, then ${\bar G}_{\varepsilon, \lambda}=[H_{cc}(G)]^-_{\varepsilon, \lambda}=[H_{cc}(G)]^-_c$ has the countable concatenation property, where $H_{cc}(G)$ denotes the countable concatenation hull of $G$, namely $H_{cc}(G)=\{\Sigma_{n=1}^{\infty}\tilde{I}_{A_n}x_n:\{x_n,n\in N\}$ is a sequence in $G$ and $\{A_n,n\in N\}$ is a countable partition of $\Omega$ to $\mathcal{F}\}$. \end{theorem} \begin{proof} (1). $\bar{G}_c=\bar{G}_{\varepsilon,\lambda}$ is by Proposition 2.27, and thus we only need to prove that $\bar{G}_{\varepsilon,\lambda}$ has the countable concatenation property. Let $\{x_n,n\in N\}$ be a given sequence in $\bar{G}_{\varepsilon,\lambda}$ and $\{A_n,n\in N\}$ a countable partition of $\Omega$ to $\mathcal{F}$, then by the countable concatenation property of $E$ there is $x^{\ast}\in E$ such that $x^{\ast}=\Sigma_{n=1}^{\infty}\tilde{I}_{A_n}x_n$. We claim that $x^{\ast}\in \bar{G}_{\varepsilon,\lambda}$, namely, $(x^{\ast}+N_{\theta}(\mathcal{Q},\varepsilon^{\ast},\lambda^{\ast}))\bigcap G\neq \emptyset$ for any given $\varepsilon^{\ast}>0,\lambda^{\ast}>0$ with $0<\lambda^{\ast}<1$ and any finite subset $\mathcal{Q}$ of $\mathcal{P}$, where $N_{\theta}(\mathcal{Q},\varepsilon^{\ast},\lambda^{\ast})=\{x\in E~|~P\{\omega\in\Omega~|~\|x\|_{\mathcal{Q}}(\omega)<\varepsilon^{\ast}\}>1-\lambda^{\ast}\}$. In fact, it is clear that there exists $\bar{x}_n$ for each $x_n\in\bar{G}_{\varepsilon,\lambda}$ such that $P\{\omega\in\Omega~|~\|x_n-\bar{x}_n\|_{\mathcal{Q}}(\omega)<\varepsilon^{\ast}\}>1-\frac{1}{2^{n+1}}\lambda^{\ast}$. By the countable concatenation property of $G$, there is $\bar{x}\in G$ such that $\bar{x}=\Sigma_{n=1}^{\infty}\tilde{I}_{A_n}\bar{x}_n$. Then $P\{\omega\in\Omega~|~\|x^{\ast}-\bar{x}\|_{\mathcal{Q}}(\omega)\geq\varepsilon^{\ast}\}=\Sigma_{n=1}^{\infty}P\{\omega\in A_n~|~\|x_n-\bar{x}_n\|_{\mathcal{Q}}(\omega)\geq\varepsilon^{\ast}\}\leq\Sigma_{n=1}^{\infty}P\{\omega\in \Omega~|~\|x_n-\bar{x}_n\|_{\mathcal{Q}}(\omega)\geq\varepsilon^{\ast}\}\leq\Sigma_{n=1}^{\infty}\frac{1}{2^{n+1}}\lambda^{\ast}=\frac{1}{2}\lambda^{\ast}$, namely $P\{\omega\in \Omega~|~\|x^{\ast}-\bar{x}\|_{\mathcal{Q}}(\omega)<\varepsilon^{\ast}\}\geq1-\frac{1}{2}\lambda^{\ast}>1-\lambda^{\ast}$, that is to say, $\bar{x}\in(x^{\ast}+N_{\theta}(\mathcal{Q},\varepsilon^{\ast},\lambda^{\ast}))\bigcap G$. (2). By (1), $[H_{cc}(G)]^{-}_{c}=[H_{cc}(G)]^{-}_{\varepsilon,\lambda}$ has the countable concatenation property. Thus we only need to prove that $\bar{G}_{\varepsilon,\lambda}=[H_{cc}(G)]^{-}_{\varepsilon,\lambda}$. We can suppose , without loss of generality, that $\theta\in G$. Then for any $x=\Sigma_{n=1}^{\infty}\tilde{I}_{A_i}g_i\in H_{cc}(G)$, it is obvious that $\{\Sigma_{i=1}^{n}\tilde{I}_{A_i}g_i~|~n\in N\}$ is a $\mathcal{T}_{\varepsilon,\lambda}$--cauchy sequence in $G$ convergent to $x$ since $\{A_n,n\in N\}$ is a countable partition of $\Omega$ to $\mathcal{F}$, which means that $x\in \bar{G}_{\varepsilon,\lambda}$, namely $[H_{cc}(G)]^{-}_{\varepsilon,\lambda}\subset \bar{G}_{\varepsilon,\lambda}$, so $[H_{cc}(G)]^{-}_{\varepsilon,\lambda}=\bar{G}_{\varepsilon,\lambda}$. \hfill $\square$ \end{proof} \begin{remark} We can illustrate that (1) may be not true if $E$ does not have the countable concatenation property, such an example is omitted to save space. \end{remark} \subsection{ On the precise relation between random conjugate spaces under the two kinds of topologies} For any $RLC$ module $(E,\mathcal{P})$, Guo already pointed out that $E^{\ast}_{\varepsilon,\lambda}$ has the countable concatenation property, cf.\cite{TXG-JFA}, let $H_{cc}(E^{\ast}_{c})$ be the countable concatenation hull of $E^{\ast}_{c}$ in $E^{\ast}_{\varepsilon,\lambda}$. The main result of this subsection is the following: \begin{theorem} $E^{\ast}_{\varepsilon,\lambda}:=(E,\mathcal{P})^{\ast}_{\varepsilon,\lambda}=(E,\mathcal{P}_{cc})^{\ast}_{c}=H_{cc}(E^{\ast}_{c})$, where, please recall $E^{\ast}_{c}:=(E,\mathcal{P})^{\ast}_{c}$. \end{theorem} To prove Theorem 2.30 and meet the needs of Subsection 3.3.1, we first recall from \cite{TXG-XXC} Lemma 2.31 below and the two corollaries to it. \begin{lemma}($See$ \cite{TXG-XXC}.) Let $X$ be a linear space over $K$, $\{p_n:X\rightarrow L^{0}_{+}(\mathcal{F})~|~n\in N\}$ a sequence of random seminorms on $X$ and $f:X\rightarrow L^{0}(\mathcal{F},K)$ a random linear functional such that $\Sigma_{n=1}^{\infty}p_n(x)$ converges in probability for each $x\in X$ and $|f(x)|\leq\Sigma_{n=1}^{\infty}p_n(x)$ for each $x\in X$. Then there is a sequence $\{f_n~|~n\in N\}$ of random linear functionals such that \noindent (1). $|f_n(x)|\leq p_n(x)$ for all $x\in X$ and $n\in N$; \noindent (2). $f(x)=\Sigma_{n=1}^{\infty}f_n(x)$ for all $x\in X$. \end{lemma} \begin{corollary}($See$ \cite{TXG-XXC}.) Let $X$ be an $L^{0}(\mathcal{F},K)$--module, $f:X\rightarrow L^{0}(\mathcal{F},K)$ an $L^{0}(\mathcal{F},K)$--linear function, $\{p_n:X\rightarrow L^{0}_{+}(\mathcal{F})~|~n\in N\}$ a sequence of $L^0$--seminorms on $X$ and $\{A_n,n\in N\}$ a countable partition of $\Omega$ to $\mathcal{F}$ such that $|f(x)|\leq\Sigma_{n=1}^{\infty}\tilde{I}_{A_n}p_n(x)$ for all $x\in X$. Then there is a sequence $\{f_n:n\in N\}$ of $L^{0}(\mathcal{F},K)$--linear functions such that \noindent (1). $|f_n(x)|\leq p_n(x)$ for all $x\in X$ and $n\in N$; \noindent (2). $f(x)=\Sigma_{n=1}^{\infty}\tilde{I}_{A_n}(f_n(x))$ for all $x\in X$. \end{corollary} \begin{corollary} ($See$ \cite{TXG-XXC}.) Let $X$ be an $L^{0}(\mathcal{F},K)$--module, $f:X\rightarrow L^{0}(\mathcal{F},K)$ an $L^{0}(\mathcal{F},K)$--linear function and $\{p_i:X\rightarrow L^{0}_+(\mathcal{F})~|~1\leq i\leq n\}$ n $L^{0}$--seminorms such that $|f(x)|\leq\Sigma_{i=1}^{n}p_i(x)$ for all $x\in X$. Then for each $1\leq i\leq n$ there is an $L^{0}(\mathcal{F},K)$--linear function $f_i$ such that \noindent (1). $|f_i(x)|\leq p_i(x)$ for all $x\in X$ and $1\leq i\leq n$; \noindent (2). $f(x)=\Sigma_{i=1}^{n}f_n(x)$ for all $x\in X$. \end{corollary} We can now prove Theorem 2.30. \noindent {Proof of Theorem 2.30.} Since $\mathcal{P}$ and $\mathcal{P}_{cc}$ induces the same $(\varepsilon,\lambda)$--topology on $E$, $E^{\ast}_{\varepsilon,\lambda}:=(E,\mathcal{P})^{\ast}_{\varepsilon,\lambda}=(E,\mathcal{P}_{cc})^{\ast}_{\varepsilon,\lambda}=(E,\mathcal{P}_{cc})^{\ast}_{c}$, where the last equality comes from the countable concatenation property of $\mathcal{P}_{cc}$ by Corollary 2.21. It remains to prove that $(E,\mathcal{P}_{cc})^{\ast}_{c}=H_{cc}(E^{\ast}_c)$ and we only needs to check that $(E,\mathcal{P}_{cc})^{\ast}_{c}\subset H_{cc}(E^{\ast}_c)$. Let $f$ be any element of $(E,\mathcal{P}_{cc})^{\ast}_{c}$. Since $\mathcal{P}_{cc}$ is invariant under the operation of finitely many suprema, then there are $\|\cdot\|\in \mathcal{P}_{cc}$ and $\xi\in L^{0}_{++}(\mathcal{F})$ such that $|f(x)|\leq\xi\|x\|$ for all $x\in E$. Let $\|\cdot\|=\Sigma_{n=1}^{\infty}\tilde{I}_{A_n}\|\cdot\|_{\mathcal{Q}_n}$, where $\{A_n,n\in N\}$ is some countable partition of $\Omega$ to $\mathcal{F}$ and each $\mathcal{Q}_n\in \mathcal{P}_f$, then by Corollary 2.32 there is a sequence $\{f_n,n\in N\}$ of $L^{0}(\mathcal{F},K)$--linear functions such that \noindent (1). $|f_n(x)|\leq\xi\|x\|_{\mathcal{Q}_n}$ for all $x\in E$ and $n\in N$; \noindent (2). $f(x)=\Sigma_{n=1}^{\infty}\tilde{I}_{A_n}(f_n(x))$ for all $x\in E$. (1) shows that each $f_n\in E^{\ast}_c$ and (2) further shows that $f=\Sigma_{n=1}^{\infty}\tilde{I}_{A_n}f_n$, so $f\in H_{cc}(E^{\ast}_c)$. \hfill $\square$ For the further study of random conjugate spaces, please refer to \cite{TXG-SEZ}. \subsection{Bounded sets under the locally $L^0-$convex topology} Let $(E,\mathcal{T})$ be a locally $L^{0}$--convex $L^{0}(\mathcal{F},K)$--module. $A\subset E$ is said to be $\mathcal{T}$--bounded if $A$ can be $L^0$--absorbed by every neighborhood of $\theta$. The main results in this subsection are Theorems 2.34 and 2.35 below. \begin{theorem} $(${\bf Random resonance theorem}$)$ Let $(E_1,\|\cdot\|_1)$ and $(E_2,\|\cdot\|_2)$ be two $RN$ modules such that $E_1$ is $\mathcal{T}_c$--complete and has the countable concatenation property. For a subset $\{T_{\alpha},\alpha\in\Lambda\}$ of $B(E_1,E_2)$, then $\{T_{\alpha},\alpha\in\Lambda\}$ is $\mathcal{T}_c$--bounded in $B(E_1,E_2)$ iff $\{T_{\alpha}x,\alpha\in\Lambda\}$ is $\mathcal{T}_c$--bounded in $E_2$ for all $x\in E_1$. \end{theorem} Let $(E_1,\|\cdot\|)$ and $(E_2,\|\cdot\|)$ be $RN$ modules over $K$ with base $(\Omega,\mathcal{F},P)$. It is easy to prove that a linear operator $T:E_1\rightarrow E_2$ belongs to $B(E_1,E_2)$ iff $T$ is a continuous module homomorphism from $(E_1,\mathcal{T}_c)$ to $(E_2,\mathcal{T}_c)$. Hence Theorem 2.34 also gives a resonance theorem for a family of continuous module homomorphisms from $(E_1,\mathcal{T}_c)$ to $(E_2,\mathcal{T}_c)$. \begin{theorem} Let $(E,\mathcal{P})$ be an $RLC$ module over $K$ with base $(\Omega,\mathcal{F},P)$ and $A\subset E$. Then $A$ is $\mathcal{T}_c$--bounded iff $f(A)$ is $\mathcal{T}_c$--bounded in $(L^{0}(\mathcal{F},K),\mathcal{T}_c)$ for every $f\in E^{\ast}_c$. \end{theorem} Theorems 2.34 and 2.35 can be implied by the work on random resonance theorem at the earlier stage of $RLC$ spaces. To see this, let us recall: \begin{definition} (See \cite{TXG-PHD,TXG-Module}.) Let $(E,\mathcal{P})$ be an $RLC$ space over $K$ with base $(\Omega,\mathcal{F},P)$. A set $A\subset E$ is said to be a.s. bounded if $\bigvee\{\|a\|:a\in A\}\in L^{0}_{+}(\mathcal{F})$ for each $\|\cdot\|\in\mathcal{P}$. \end{definition} Lemma 2.37 below is clear by definition. \begin{lemma} Let $(E,\mathcal{P})$ be an $RLC$ module. Then a set $A$ of $E$ is $\mathcal{T}_c$--bounded iff $A$ is a.s. bounded. \end{lemma} Thus, by Proposition 2.26 one can easily see that Theorem 2.34 is a corollary to (2) of Proposition 2.38 below. Since a $\mathcal{T}_{\varepsilon,\lambda}$--complete $RN$ module is a Frech$\acute{e}$t space (namely a complete metrizable linear topological space), (1) of Proposition 2.38 can be proved with the aid of the classical resonance theorem (see \cite{Yosida}). Hence (2) of Proposition 2.38 is more interesting. \begin{proposition} ($See$ \cite{TXG-PHD,TXG-Module}.) Let $(E_1,\|\cdot\|_1)$ and $(E_2,\|\cdot\|_2)$ be two $RN$ modules over $K$ with base $(\Omega,\mathcal{F},P)$ such that $E_1$ is $\mathcal{T}_{\varepsilon,\lambda}$--complete. Given a subset $\{T_{\alpha},\alpha\in\Lambda\}$ in $B(E_1,E_2)$, we have the following: \noindent (1). $\{T_{\alpha},\alpha\in\Lambda\}$ is $\mathcal{T}_{\varepsilon,\lambda}$--bounded in $B(E_1,E_2)$ iff $\{T_{\alpha}x,\alpha\in\Lambda\}$ is $\mathcal{T}_{\varepsilon,\lambda}$--bounded in $E_2$ for all $x\in E_1$. \noindent (2). $\{T_{\alpha},\alpha\in\Lambda\}$ is a.s. bounded in $B(E_1,E_2)$ iff $\{T_{\alpha}x,\alpha\in\Lambda\}$ is a.s. bounded in $E_2$ for all $x\in E_1$. \end{proposition} One can also see that Theorem 2.35 is only a corollary to (3) of Proposition 2.39 below. However, (1) and (2) of Proposition 2.39 are also interesting because of the use of random conjugate spaces. \begin{proposition} ($See$ \cite{TXG-PHD,TXG-Module,TXG-Sur}.) Let $(E,\mathcal{P})$ be an $RLC$ space over $K$ with base $(\Omega,\mathcal{F},P)$ and $A$ a subset of $E$. Then we have: \noindent (1). $A$ is $\mathcal{T}_{\varepsilon,\lambda}$--bounded iff $f(A)$ is $\mathcal{T}_{\varepsilon,\lambda}$--bounded in $(L^0(\mathcal{F},K),\mathcal{T}_{\varepsilon,\lambda})$ for each $f\in E^{\ast}_{I}$. \noindent (2). $A$ is $\mathcal{T}_{\varepsilon,\lambda}$--bounded iff $f(A)$ is $\mathcal{T}_{\varepsilon,\lambda}$--bounded in $(L^0(\mathcal{F},K),\mathcal{T}_{\varepsilon,\lambda})$ for each $f\in E^{\ast}_{II}$. \noindent (3). $A$ is a.s. bounded iff $f(A)$ is a.s. bounded in $(L^0(\mathcal{F},K),|\cdot|)$ for each $f\in E^{\ast}_{I}$. \noindent (4). $A$ is a.s. bounded iff $f(A)$ is a.s. bounded in $(L^0(\mathcal{F},K),|\cdot|)$ for each $f\in E^{\ast}_{II}$. \end{proposition} \begin{remark} $\mathcal{T}_c$--bounded sets (or more general a.s. bounded sets) have played a crucial role in random duality theory in this paper and in \cite{TXG-XXC}, whereas $\mathcal{T}_{\varepsilon,\lambda}$--bounded sets are important in other fields, for example, they are called ``probabilistically bounded sets" in the theory of probabilistic normed spaces (see \cite{SS}) and are called ``stochastically bounded sets" for Banach space-valued random elements in probability theory in Banach spaces. \end{remark} \section{Random convex analysis over random locally convex modules under the locally $L^{0}$--convex topology} As has been mentioned in Section 1 of this paper, Lemma 2.28 of \cite{FKV} is key but false, we have to first correct it and the closely related basic results for the further development of random convex analysis. The first two parts of this section not only have corrected them but also have refreshed the related basic results based on the precise relation between $E^{\ast}_{\varepsilon,\lambda}$ and $E^{\ast}_{c}$. \subsection{Separation under the locally $L^0$--convex topology} First, Lemma 2.28 of \cite{FKV} should be modified as Lemma 3.1 below. \begin{lemma} Let $(E,\mathcal{P})$ be an $RLC$ module over $K$ with base $(\Omega,\mathcal{F},P)$ such that $\mathcal{P}$ has the countable concatenation property, $M$ a $\mathcal{T}_c$--closed subset with the countable concatenation property and $x\in E$ such that $\tilde{I}_A\{x\}\bigcap\tilde{I}_A M=\emptyset$ for all $A\in \mathcal{F}$ with $P(A)>0$. Then there is an $L^0$--convex, $L^0$--absorbent and $L^0$--balanced $\mathcal{T}_c$--neighborhood $U$ of $\theta$ such that $$\tilde{I}_A(x+U)\bigcap\tilde{I}_A(M+U)=\emptyset$$ for all $A\in\mathcal{F}$ with $P(A)>0$. \end{lemma} Lemma 2.28 of \cite{FKV} only requires $M$ to satisfy the condition that $\tilde{I}_{A}M+\tilde{I}_{A^c}\subset M$ for all $A\in\mathcal{F}$, which is much weaker than the countable concatenation property as assumed in our Lemma 3.1. Example 3.2 below is a counterexample to Lemma 2.28 of \cite{FKV}. Let us first point out that Lemma 3.1 has been implied by Lemma 3.10 of \cite{TXG-JFA}, which can be explained as follows. Let us recalled from \cite{TXG-JFA}: for each $\mathcal{Q}\in\mathcal{P}_f$ and $\varepsilon\in L^{0}_{++}(\mathcal{F})$, let $U_{\mathcal{Q},\varepsilon}[x]=\{y\in E~|~\|x-y\|_{\mathcal{Q}}\leq\varepsilon\}$, $e^{\ast}_{\mathcal{Q}}(x,M)=\bigwedge\{\varepsilon\in L^0_{++}(\mathcal{F})~|~U_{\mathcal{Q},\varepsilon}[x]\bigcap M\neq\emptyset\}$ and $e^{\ast}(x,M)=\bigvee\{e^{\ast}_{\mathcal{Q}}(x,M)~|~\mathcal{Q}\in\mathcal{P}_{f}\}$. Then Lemma 3.10 of \cite{TXG-JFA} shows that $e^{\ast}(x,M)\bigwedge 1\in L^0_{++}(\mathcal{F})$, which is just what the proof of Lemma 2.28 of \cite{FKV} needs, the remainder of the proof of Lemma 3.1 is the same as the corresponding part of the proof of Lemma 2.28 of \cite{FKV}. \begin{example} Let $(\Omega,\mathcal{F},P)$ be a nonatomic probability space (namely $\mathcal{F}$ does not include any $P$--atoms), $(E,\mathcal{P})=(L^{0}(\mathcal{F},R),|\cdot|)$ and $M=\{x\in E~|~$there exists a positive number $m_x$ such that $x>m_x$ on $\Omega\}$. Then Claim 3.3 below shows that $M$ is $L^0$--convex, $\mathcal{T}_c$--closed and $\mathcal{T}_c$--open. Further, Claim 3.4 below shows that $\tilde{I}_A\{0\}\bigcap\tilde{I}_A M=\emptyset$ for all $A\in\mathcal{F}$ with $P(A)>0$, but for each $L^0$--convex, $L^0$--absorbent and $L^0$--balanced $\mathcal{T}_c$--neighborhood $U$ of 0 there is an $A_U\in\mathcal{F}$ with $P(A_U)>0$ such that $$\tilde{I}_{A_U}U\bigcap\tilde{I}_{A_U}(M+U)\neq\emptyset.$$Thus this provides a counterexample to Lemma 2.28 of \cite{FKV}. \end{example} \begin{claim} $M$ in Example 3.2 is $L^0$--convex, $\mathcal{T}_c$--closed and $\mathcal{T}_c$--open. \end{claim} \begin{proof} First, it is obvious that $M$ is $L^0$--convex. Second, $M$ is $\mathcal{T}_c$--open. For any $y\in M$, by definition there is some positive number $m_y$ such that $y>m_y$ on $\Omega$. Let $\varepsilon^0\equiv\frac{1}{2}m_y$ and $\varepsilon$ be the equivalence class of $\varepsilon^0$, then $\varepsilon\in L^0_{++}(\mathcal{F})$ and hence $B(\varepsilon):=\{x\in E~|~|x|\leq\varepsilon\}$ is a $\mathcal{T}_c$--neighborhood of 0, it is also easy to check that $y+B(\varepsilon)\subset M$. Finally, $M$ is also $\mathcal{T}_c$--closed, namely $E\setminus M$ is $\mathcal{T}_c$--open, which will be proved in the following three cases. Case (1): when $y\in E\setminus M$ and $y\not\in L^0_+(\mathcal{F})$, there is $D\in\mathcal{F}$ with $P(D)>0$ such that $y<0$ on $D$. Let $\varepsilon=\tilde{I}_{D^c}+\frac{1}{2}\tilde{I}_{D}|y|(\in L^{0}_{++}(\mathcal{F}))$ and $B(\varepsilon)=\{x\in E~|~|x|\leq\varepsilon\}$, then $y+B(\varepsilon)\subset E\setminus M$. In fact, for any $z\in y+B(\varepsilon)$, $z-y\leq \tilde{I}_{D^c}+\frac{1}{2}\tilde{I}_{D}|y|$ implies that $z\leq y+\frac{1}{2}|y|=-\frac{1}{2}|y|<0$ on $D$, namely $z\in E\setminus M$. Case (2): when $y\in E\setminus M$, $y\in L^0_+(\mathcal{F})$ and $y\not\in L^{0}_{++}(\mathcal{F})$, then there is $D\in\mathcal{F}$ with $P(D)>0$ such that $y=0$ on $D$. Since $(\Omega,\mathcal{F},P)$ is nonatomic, there is a countable partition $\{D_n,n\in N\}$ of $D$ to $\mathcal{F}$ such that $P(D_n)=\frac{1}{2^n}P(D)$ for each $n\in N$. Let $\varepsilon=\tilde{I}_{D^c}+\Sigma_{n=1}^{\infty}\frac{1}{n}\tilde{I}_{D_n} (\in L^{0}_{++}(\mathcal{F}))$ and $B(\varepsilon)=\{x\in E~|~|x|\leq\varepsilon\}$, then $z\leq \frac{1}{n}$ on $D_n$ for any $z\in y+B(\varepsilon)$, which implies that $P\{\omega\in\Omega~|~z(\omega)\leq\frac{1}{n}\}\geq P(D_n)>0$ for all $n\in N$, namely $y+B(\varepsilon)\subset E\setminus M$. Case (3): when $y\in E\setminus M$ and $y\in L^0_{++}(\mathcal{F})$, then $P\{\omega\in \Omega~|~y(\omega)<\frac{1}{n}\}>0$ for each $n\in N$ by the definition of $M$. Let $H_n=[y<\frac{1}{n}]$ and $D_n=[\frac{1}{n+1}\leq y<\frac{1}{n}]$ for any $n\in N$, then $D_i\bigcap D_j=\emptyset$ for $i\neq j$ and $H_n=\Sigma_{i=n}^{\infty}D_i$. Obviously, it is impossible that there is some $k\in N$ such that $P(D_n)=0$ for all $n\geq k$. So, we can suppose, without loss of generality, that $P(D_n)>0$ for each $n\in N$. Let $D=\Sigma_{n=1}^{\infty}D_n$, $\varepsilon=I_{D^{c}}+\Sigma_{n=1}^{\infty} \frac{1}{n}I_{D_{n}}(\in L^{0}_{++}(\mathcal{F}))$ and $B(\varepsilon) =\{x\in E|~|x|\leq \varepsilon\}$, then for any $z\in y+B(\varepsilon)$, $z\leq\frac{2}{n}$ on $D_n$, which means that $P\{\omega\in \Omega|z(\omega)\leq\frac{2}{n}\}\geq P(D_n)>0$ for each $n\in N$, and hence $z\in E\setminus M$. \hfill $\square$ \end{proof} \begin{claim} Let $(E, {\mathcal P})$ and $M$ be the same as in Example 3.2. Then ${\tilde I}_A\{0\}\cap {\tilde I}_AM=\emptyset$ for all $A\in {\mathcal F}$ with $P(A)>0$. But for any $L^0$--convex, $L^0$--absorbent and $L^0$--balanced ${\mathcal T}_c$--neighborhood $U$ of $0$ there is always $A_U\in {\mathcal F}$ with $P(A_U)>0$ such that ${\tilde I}_{A_U}U\cap {\tilde I}_{A_U}(M+U)\neq\emptyset$. \end{claim} \begin{proof} There is $\varepsilon \in L^0_{++}({\mathcal F})$ for $U$ stated above such that $B(\varepsilon):=\{x\in E~|~|x|\leq\varepsilon\}\subset U$. For a representative $\varepsilon^0$ of $\varepsilon$, let $A_1=\{\omega\in \Omega~|~\varepsilon^0(\omega)\geq 1\}$ and $A_n=\{\omega\in\Omega~|~\frac{1}{n}\leq \varepsilon^0(\omega)<\frac{1}{n-1}\}$ for $n\geq 2$, then it is clear that $\sum^{\infty}_{n=1}P(A_n)=1$, and hence there is some $n_0\in N$ such that $P(A_{n_0})>0$. Let $A_U=A_{n_0}$ and $y_0={\tilde I}_{A^c_U}+{\tilde I}_{A_U}\varepsilon$, then $\frac{1}{n_0}\leq y_0<\frac{1}{n_0-1}$ on $A_U$ ( note: this is also true for $n_0=1$ ) and $y_0\geq \frac{1}{n_0}$ on $\Omega$ (namely, $y_0\in M$). Since ${\tilde I}_{A_U}y_0={\tilde I}_{A_U}\varepsilon\in {\tilde I}_{A_U}B(\varepsilon)\subset {\tilde I}_{A_U} U$ and ${\tilde I}_{A_U}y_0\in {\tilde I}_{A_U}M\subset{\tilde I}_{A_U}(M+U)$, so ${\tilde I}_{A_U}U\cap {\tilde I}_{A_U}(M+U)\neq\emptyset$. \hfill $\square$ \end{proof} In fact, Example 3.2 is also a counterexample to Theorem 2.8 of \cite{FKV}, which can be explained as follows. Since $(E, {\mathcal P})=(L^0({\mathcal F}, R), |\cdot|)$ is an $RN$ module, $|\cdot|$ has the countable concatenation property and $E^\ast_c=E^\ast_{\varepsilon, \lambda}$. It is obvious that $0\in {\overline{M}}_{\varepsilon, \lambda}$ ( namely, the ${\mathcal T}_{\varepsilon, \lambda}$--closure of $M$ ), and hence for each $f\in E^\ast_c=E^\ast_{\varepsilon, \lambda}$ there exists a sequence $\{y_n, n\in N\}$ in $M$ such that $\{f(y_n): n\in N\}$ converges in probability P to $0$, which means that it is impossible that there exists $f\in E^\ast_c$ such that $0=f(0)>\bigvee\{f(y): y\in M\}$ on $\Omega$. Theorem 2.8 of \cite{FKV} should be modified as follows: \begin{theorem} Let $(E, {\mathcal P})$ be an $RLC$ module over $K$ with base $(\Omega, {\mathcal F}, P)$ such that ${\mathcal P}$ has the countable concatenation property, $x\in E$ and $M$ a nonempty ${\mathcal T}_c$--closed $L^0$--convex subset with the countable concatenation property. If ${\tilde I}_A\{x\}\cap {\tilde I}_AM=\emptyset$ for all $A\in {\mathcal F}$ with $P(A)>0$, then there is $f\in E^\ast_c$ such that $$(Re f)(x)>\bigvee\{(Ref)(y): y\in M\} ~\hbox{on}~\Omega.$$ \end{theorem} Theorem 3.5 first occurred in \cite{TXG-JFA} in its current form, which can be derived from Theorem 3.6 below. Here, we give a more simpler form of Theorem 3.6 for convenience in use, this form has been implied in the process of the proof of Theorem 3.7 of \cite{TXG-JFA}. \begin{theorem} ($See$ \cite{TXG-JFA}.) Let $(E, {\mathcal P})$ be an random locally convex module over $K$ with base $(\Omega, {\mathcal F}, P)$, $x\in E$ and $M$ a ${\mathcal T}_{\varepsilon, \lambda}$--closed $L^0$--convex nonempty subset of $E$. If $x\notin M$, then there is $f\in E^\ast_{\varepsilon, \lambda}$ such that:\\ (1). $(Ref)(x)>\bigvee\{(Ref)(y): y\in M\}~\hbox{on}~A$;\\ (2). $(Ref)(x)=\bigvee\{(Ref)(y): y\in M\}~\hbox{on}~A^c$.\\ Here, $A$ is a representative of $[d^\ast(x, M)>0]$ and $d^\ast(x, M)=\bigvee\{d^\ast_{\mathcal{Q}}(x, M): \mathcal{Q}\in {\mathcal P}_f\}$, where $d^\ast_{\mathcal{Q}}(x, M)=\bigwedge\{\|x-y\|_{\mathcal{Q}}: y\in M\}$ for $\mathcal{Q}\in {\mathcal P}_f$. Since $x\not\in M$ iff $d^\ast(x, M)>0$, $P(A)>0$. \end{theorem} \begin{remark} When $(E, {\mathcal P})$ is an $RN$ module, $d^\ast(x, M)$ is just the random distance from $x$ to $M$. Thus Theorem 3.6 is best possible from the degree that $f$ separates $x$ from $M$. \end{remark} The condition that ${\tilde I}_A\{x\}\cap {\tilde I}_AM=\emptyset$ for all $A\in {\mathcal F}$ with $P(A)>0$ is to guarantee the separation of $x$ from $M$ by $f$ with probability $1$, but the condition is too strong to be easily satisfied in applications. In fact, Theorem 3.6 also yields a kind of generalization for Theorem 3.5, namely Corollary 3.8 below. \begin{corollary} Let $(E, {\mathcal P})$ be an $RLC$ module over $K$ with base $(\Omega, {\mathcal F}, P)$ such that ${\mathcal P}$ has the countable concatenation property, $x\in E$ and $M\subset E$ a nonempty ${\mathcal T}_c$--closed $L^0$--convex set with the countable concatenation property. If $x\notin M$, then there is $f\in E^\ast_c$ such that:\\ (1). $(Ref)(x)>\bigvee\{(Ref)(y): y\in M\}~\hbox{on}~A$;\\ (2). $(Ref)(x)=\bigvee\{(Ref)(y): y\in M\}~\hbox{on}~A^c$.\\ Here $A$ and $d^\ast(x, M)$ are the same as in Theorem 3.6. \end{corollary} Example 3.2 shows that it is essential that $M$ has the countable concatenation property in Corollary 3.8, but random duality theory in Subsection 3.3 needs another generalization of Corollary 3.8, namely Corollary 3.9 below, in which the condition that ${\mathcal P}$ has the countable concatenation property is removed but (1) of Corollary 3.8 only holds on a subset $B$ of $A$ with $P(B)>0$. In fact, Corollary 3.9 is more convenient for use and its proof is based on Theorem 2.30. \begin{corollary} Let $(E, {\mathcal P})$ be an $RLC$ module over $K$ with base $(\Omega, {\mathcal F}, P)$, $x\in E$ and $M\subset E$ a nonempty ${\mathcal T}_c$--closed $L^0$--convex set with the countable concatenation property. If $x\notin M$, then there exist $f\in E^\ast_c$ and $B\in {\mathcal F}$ with $P(B)>0$ such that:\\ (1). $(Ref)(x)>\bigvee\{(Ref)(y): y\in M\}~\hbox{on}~B$;\\ (2). $(Ref)(x)=\bigvee\{(Ref)(y): y\in M\}~\hbox{on}~B^c$. \end{corollary} \begin{proof} We consider the separation problem in $(E, {\mathcal P}_{cc})$. Since ${\mathcal P}_{cc}$ has the countable concatenation property and the locally $L^0$--convex topology induced by ${\mathcal P}_{cc}$ is stronger than that induced by ${\mathcal P}$. We can apply Corollary 3.8 to $(E, {\mathcal P}_{cc})$, $x$ and $M$, then there is $g\in (E, {\mathcal P}_{cc})^\ast_c$ such that:\\ (3). $(Reg)(x)>\bigvee\{(Reg)(y): y\in M\}~\hbox{on}~A$;\\ (4). $(Reg)(x)=\bigvee\{(Reg)(y): y\in M\}~\hbox{on}~A^c$.\\ Here, please note that ${\mathcal P}$ and ${\mathcal P}_{cc}$ induce the same $d^\ast(x, M)$, so $A$ is still a representative of $[d^\ast(x, M)>0]$. Since $(E, {\mathcal P}_{cc})^\ast_c=H_{cc}(E^\ast_c)$ by Theorem 2.30, $g=\sum^\infty_{n=1}{\tilde I}_{A_n}g_n$ for some countable partition $\{A_n, n\in N\}$ of $\Omega$ to ${\mathcal F}$ and some sequence $\{g_n, n\in N\}$ in $E^\ast_c$. Let $n_0\in N$ be such that $P(A\cap A_{n_0})>0$ and further let $B=A\cap A_{n_0}$ and $f={\tilde I}_{A\cap A_{n_0}}g_{n_0}$, then $f$ and $B$ meet the needs of (1) and (2). \hfill \hfill $\square$ \end{proof} In Theorem 3.6, $f$ belongs to $E^\ast_{\varepsilon, \lambda}$, but the study of random admissible topology in Subsection 3.3.2 requires an $f\in E^\ast_c$ to separate a point from a ${\mathcal T}_{\varepsilon, \lambda}$--closed $L^0$--convex subset. By noting that $E^\ast_{\varepsilon, \lambda}=H_{cc}(E^\ast_c)$, we can use the same reasoning as in the proof of Corollary 3.9 to obtain the following generalization of Theorem 3.6: \begin{corollary} Let $(E, {\mathcal P})$ be an $RLC$ module over $K$ with base $(\Omega, {\mathcal F}, P)$, $x\in E$ and $M\subset E$ a nonempty ${\mathcal T}_{\varepsilon, \lambda}$--closed $L^0$--convex subset. If $x\notin M$, then there are $f\in E^\ast_c$ and some $B\in {\mathcal F}$ with $P(B)>0$ such that:\\ (1). $(Ref)(x)>\bigvee\{(Ref)(y): y\in M\}~\hbox{on}~B$;\\ (2). $(Ref)(x)=\bigvee\{(Ref)(y): y\in M\}~\hbox{on}~B^c$. \end{corollary} \begin{remark} Let $\xi$ be any element in $L^0({\mathcal F}, K)$ and $\xi_0$ a representative of $\xi$. Define $\xi_0^{-1}\colon\Omega\to K$ by $\xi_0^{-1}(\omega)=(\xi_0(\omega))^{-1}$ if $\xi_0(\omega)\neq 0$ and by $0$ if $\xi_0(\omega)=0$, then $\xi^{-1}:=$ the equivalence class of $\xi_0^{-1}$ is called the generalized inverse of $\xi$. $|\xi|^{-1}\xi$ is called the sign of $\xi$, denoted by $sgn(\xi)$, then ${\overline {sgn(\xi)}}\xi=|\xi|$, where ${\overline {sgn(\xi)}}$ stands for the complex conjugate of $sgn(\xi)$. Further, we also have that $\xi\cdot\xi^{-1}=\xi^{-1}\cdot\xi=I_{[\xi\neq 0]}$. If $M$ in Corollary 3.9 or Corollary 3.10 is also $L^0$--balanced, then one can make use of the notion of the sign for element in $L^0({\mathcal F}, K)$ to see that (1) and (2) of the two corollaries can be rewritten as ( cf. \cite{TXG-strict} ):\\ (1). $|f(x)|>\bigvee\{|f(y)|: y\in M\}~\hbox{on}~B$;\\ (2). $|f(x)|=\bigvee\{|f(y)|: y\in M\}~\hbox{on}~B^c$. \end{remark} Let $\xi=|f(x)|$ and $\eta=\bigvee\{|f(y)|: y\in M\}$, then multiplying the above two sides by $(\frac{\xi+\eta}{2})^{-1}$ and replacing $f$ with $(\frac{\xi+\eta}{2})^{-1}f$ ( still denoted by $f$ ) will obtain the following two relations:\\ (3). $|f(x)|>\bigvee\{|f(y)|: y\in M\}$;\\ (4). $|f(x)|\nleqslant 1$ and $\bigvee\{|f(y)|: y\in M\}\leq 1$.\\ (3) and (4) will be used in the proof of random bipolar theorem in Subsection 3.3.1. All the above results from Theorem 3.5 to Corollary 3.10 are concerned with the separation between a point and a closed subset. Theorem 3.12 below, due to \cite{FKV}, is concerned with the separation between two $L^0$--convex sets with one of them open, which is peculiar to the locally $L^0$--convex topology since it is impossible to establish such a theorem under the $(\varepsilon, \lambda)$--topology. Since the Proof of Theorem 3.12 does not involve the problem of whether a locally $L^0$--convex topology can be induced by a family of $L^0$--seminorms, we still state it as in \cite{FKV}. \begin{theorem}($See$ \cite{FKV}.) Let $(E, {\mathcal T})$ be a Hausdorff locally $L^0$--convex $L^0({\mathcal F}, K)$--module, $M$ and $G$ two nonempty $L^0$--convex sets of $E$ with $G$ open. If ${\tilde I}_AM\cap {\tilde I}_AG=\emptyset$ for all $A\in {\mathcal F}$ with $P(A)>0$, then there is $f\in E^\ast_c$ such that:$$(Ref)(y)<(Ref)(z)~\hbox{on $\Omega$ for all $y\in G$ and $z\in M$}.$$ \end{theorem} Theorem 3.12 will be used in Section 3.4, see \cite{TXG-JFA,TXG-GS} for its slight generalizations. \subsection{The Fenchel-Moreau dual representation theorem under the locally $L^0$--convex topology} Let $E$ be an $L^0({\mathcal F})$--module and $f$ a function from $E$ to ${\bar L}^0({\mathcal F})$. The effective domain of $f$ is denoted by $dom(f):=\{x\in E~|~|f(x)|<+\infty~\hbox{on}~\Omega\}$ and the epigraph of $f$ by $eip(f):=\{(x,r)\in E\times L^0({\mathcal F})~|~f(x)\leq r\}$. $f$ is proper if $dom(f)\neq\emptyset$ and $f(x)>-\infty~\hbox{on}~\Omega$. $f$ is $L^0({\mathcal F})$--convex if $f(\xi x+(1-\xi)y)\leq \xi f(x)+(1-\xi)f(y)$ for all $x,~y\in E$ and $\xi\in L^0_+({\mathcal F})$ with $0\leq \xi\leq 1$, where the following convention is adopted: $0\cdot(\pm\infty)=0$ and $+\infty\pm(\pm\infty)=+\infty$. If $(E, {\mathcal T})$ is a topological $L^0({\mathcal F})$--module, a proper function $f$ is lower semicontinuous (or ${\mathcal T}$--lower semicontinuous if there is a possible confusion) if $\{x\in E~|~f(x)\leq r\}$ is closed for all $r\in L^0({\mathcal F})$. We can now state the main result of this subsection as Theorem 3.13 below, which is a modification and improvement of Theorem 3.8 of \cite{FKV}. \begin{theorem} Let $(E, {\mathcal P})$ be an $RLC$ module over $R$ with base $(\Omega, {\mathcal F}, P)$ such that $E$ has the countable concatenation property. If $f$ is a proper, ${\mathcal T}_c$--lower semicontinuous $L^0({\mathcal F})$--convex function from $E$ to ${\bar L}^0({\mathcal F})$, then $f^{\ast\ast}_{c}=f$. Here $f^\ast_c\colon E^\ast_c\to {\bar L}^0({\mathcal F})$ is defined by $f^\ast_c(g)=\bigvee\{g(x)-f(x)~|~x\in E\}$ for all $g\in E^\ast_c$, called the ${\mathcal T}_c$--conjugate ( or penalty ) function of $f$, and $f^{\ast\ast}_{c}\colon E \to {\bar L}^0({\mathcal F})$ by $f^{\ast\ast}_{c}(x)=\bigvee\{g(x)-f^{\ast}_{c}(g)~|~g\in E^\ast_c\}$ for all $x\in E$, called the ${\mathcal T}_c$--biconjugate function of $f$. \end{theorem} As compared with Theorem 3.8 of \cite{FKV}, Theorem 3.13 requires the additional condition that $E$ has the countable concatenation property and remove the condition that ${\mathcal P}$ has the countable concatenation property. From the sequel of this subsection, one can immediately see that the additional condition is essential, while Theorem 2.30 can be used to remove the condition on ${\mathcal P}$. To prove Theorem 3.13, let us first study the properties of an $L^0$--convex function. Let $E$ be an $L^0({\mathcal F})$--module. $f\colon E\to {\bar L}^0({\mathcal F})$ is said to be local ( or, to have the local property ) if ${\tilde I}_Af(x)={\tilde I}_Af({\tilde I}_Ax)$ for all $x\in E$ and $A\in {\mathcal F}$. It is clear that ${\tilde I}_Af(x)=f({\tilde I}_Ax)$ for all $x\in E$ and $A\in {\mathcal F}$ iff $f$ is local, if $f(0)=0$. \begin{lemma}($See$ \cite{FKV,FKV-appro}.) Let $E$ be an $L^0({\mathcal F})$--module. Then a function $f\colon E\to {\bar L}^0({\mathcal F})$ is $L^0$--convex iff $f$ is local and $eip(f)$ is $L^0$--convex. \end{lemma} According to Lemma 3.1, Proposition 3.4 and Lemma 3.10 of \cite{FKV} together can be modified to Proposition 3.15 below. \begin{proposition} Let $(E, {\mathcal P})$ be an $RLC$ module over $R$ with base $(\Omega, {\mathcal F}, P)$ such that both $E$ and ${\mathcal P}$ have the countable concatenation property. If $f\colon E\to {\bar L}^0({\mathcal F})$ is a proper and local function, then the following are equivalent:\\ (1). $f$ is ${\mathcal T}_c$--lower semicontinuous.\\ (2). $eip(f)$ is closed in $(E, {\mathcal T}_c)\times (L^0({\mathcal F}), {\mathcal T}_c)$.\\ (3). $\varliminf_\alpha f(x_\alpha)\geq f(x)$ for all the nets $\{x_\alpha, \alpha\in \Gamma\}\subset E$ with $x_\alpha\to x$ with respect to ${\mathcal T}_c$ for some $x\in E$, where $\varliminf_\alpha f(x_\alpha)=\bigvee_{\beta\in\Gamma}(\bigwedge_{\alpha\geq\beta} f(x_\alpha))$. \end{proposition} \begin{proof} It is clear that (3)$\Rightarrow$(2)$\Rightarrow$(1). As to the converse implications, one can only needs to notice that both the set $\{x\in E|f(x)\leq r\}$ for any given $r\in L^0({\mathcal F})$ and $eip(f)$ have the countable concatenation property if $E$ has the property. The remainder of proof is the same as the corresponding part of proofs of Proposition 3.4 and Lemma 3.10 of \cite{FKV} since the arguments in \cite{FKV} are feasible in the case when $E$ has the countable concatenation property (in fact, only in the case Lemma 3.1 can be applied to the sets $\{x\in E|f(x)\leq r\}$ for any given $r\in L^0({\mathcal F})$ and $eip(f)$). \hfill \hfill $\square$ \end{proof} Let $E$ be an $L^{0}(\mathcal{F},K)$--module with the countable concatenation property and $G\subset E$ a nonempty subset. $H_{cc}(G)$ always denotes the countable concatenation hull of $G$ in $E$, namely $H_{cc}(G)=\{\Sigma_{n=1}^{\infty}\tilde{I}_{A_n}g_n~|~\{A_n,n\in N\}$ is a countable partition of $\Omega$ to $\mathcal{F}$ and $\{g_n,n\in N\}$ is sequence in $G\}$. For any $x\in H_{cc}(G)$, $\Sigma_{n=1}^{\infty}\tilde{I}_{A_n}g_n$ is called a canonical representation of $x$ if $\{A_n,n\in N\}$ is a countable partition of $\Omega$ to $\mathcal{F}$ and $\{g_n,n\in N\}$ is a sequence in $G$ such that $x=\Sigma_{n=1}^{\infty}\tilde{I}_{A_n}g_n$. Lemma 3.16 below is almost obvious but frequently used in the proofs of the subsequent key theorems in Section 4 as well as in Theorem 3.13, thus we summarize and prove it as follows: \begin{lemma} Let $E$ be an $L^{0}(\mathcal{F})$--module with the countable concatenation property. Then we have the following statements: \noindent $(1)$. Let $f:E\rightarrow \bar{L}^{0}(\mathcal{F})$ have the local property and $x=\Sigma_{n=1}^{\infty}\tilde{I}_{A_n}x_n$ for some countable partition $\{A_n,n\in N\}$ of $\Omega$ to $\mathcal{F}$ and some sequence $\{x_n,n\in N\}$ in $E$, then $f(x)=\Sigma_{n=1}^{\infty}\tilde{I}_{A_n}f(x_n)$. \noindent $(2)$. Let $f:E\rightarrow \bar{L}^{0}(\mathcal{F})$ have the local property and $G\subset E$ be a nonempty subset, then $\bigvee\{f(x)~|~x\in G\}=\bigvee\{f(x)~|~x\in H_{cc}(G)\}$. \noindent $(3)$. Let $f$ and $g$ be any two functions from $E$ to $\bar{L}^{0}(\mathcal{F})$ such that they both have the local property and $G\subset E$ a nonempty subset. If $f(x)=g(x)$ for all $x\in G$, then $f(x)=g(x)$ for all $x\in H_{cc}(G)$. \noindent $(4)$. Let $\{f_{\alpha},\alpha\in\Gamma\}$ be a family of functions from $E$ to $\bar{L}^{0}(\mathcal{F})$ such that each $f_{\alpha}$ has the locally property, then $f:E\rightarrow \bar{L}^{0}(\mathcal{F})$ defined by $f(x)=\bigvee\{f_{\alpha}(x)~|~\alpha\in\Gamma\}$ for all $x\in E$, also has the local property. \end{lemma} \begin{proof} (1). $f(x)=(\Sigma_{n=1}^{\infty}\tilde{I}_{A_n})f(x)=\Sigma_{n=1}^{\infty}\tilde{I}_{A_n}f(x)=\Sigma_{n=1}^{\infty}\tilde{I}_{A_n}f(\tilde{I}_{A_n}x)= \Sigma_{n=1}^{\infty}\tilde{I}_{A_n}f(\tilde{I}_{A_n}x_n)=\Sigma_{n=1}^{\infty}\tilde{I}_{A_n}f(x_n)$. (2). Let $\xi=\bigvee\{f(x)~|~x\in G\}$ and $\eta=\bigvee\{f(x)~|~x\in H_{cc}(G)\}$, then $\xi\leq\eta$ is clear, it remains to prove $\eta\leq\xi$. For any $x\in H_{cc}(G)$, let $\Sigma_{n=1}^{\infty}\tilde{I}_{A_n}g_n$ be a canonical representation of $x$, then $f(x)=\Sigma_{n=1}^{\infty}\tilde{I}_{A_n}f(g_n)\leq\xi$, so $\eta\leq\xi$. (3). It is clear by (1). (4). It is also clear by definition. \hfill \hfill $\square$ \end{proof} We can now prove Theorem 3.13. \newproof{pot}{Proof of Theorem 3.13} \begin{pot} We first consider the special case when ${\mathcal P}$ has the countable concatenation property. Since $f$ is $L^0$--convex, $f$ is local and $eip(f)$ is $L^0$--convex by Lemma 3.15. Further, since $E$ and $L^0({\mathcal F})$ have the countable concatenation property, $eip(f)$ also has the property. To sum up, $eip(f)$ is a ${\mathcal T}_c$--closed $L^0$--convex set with the countable concatenation property in the random locally convex module $(E\times L^0({\mathcal F}), {\tilde {\mathcal P}})$, where ${\tilde {\mathcal P}}=\{\|\cdot\|+|\cdot|:\|\cdot\|\in {\mathcal P}\}$ and $(\|\cdot\|+|\cdot|)(x, r)=\|x\|+|r|$ for all $(x, r)\in E\times L^0({\mathcal F})$ and $\|\cdot\|\in {\mathcal P}$, it is also obvious that ${\tilde {\mathcal P}}$ has the countable concatenation property and $(E\times L^0({\mathcal F}), {\tilde {\mathcal P}})^\ast_c=E^\ast_c\times L^0({\mathcal F})$. By applying Theorem 3.6 of this paper rather than Theorem 2.8 of \cite{FKV} to $E\times L^0({\mathcal F})$ and $eip(f)$, one can complete the proof of the special case along the idea of proof of Theorem 3.8 of \cite{FKV}. Now, we consider the general case, namely ${\mathcal P}$ may not necessarily have the countable concatenation property. In fact, based on Theorem 2.30, it is easy to complete the proof of the general case. We consider the problem in $(E, {\mathcal P}_{cc})$. Since ${\mathcal P}_{cc}$ has the countable concatenation property and the locally $L^0$--convex topology induced by ${\mathcal P}_{cc}$ is stronger than that induced by ${\mathcal P}$, applying what has been proved above to $f$ and $(E, {\mathcal P}_{cc})$ we can obtain:$$f(x)=\bigvee\{u(x)-f^\ast_c(u)~|~u\in (E, {\mathcal P}_{cc})^\ast_c\}~\hbox{ for all $x\in E$.}$$ Since $f^\ast_c$ has the local property and $u(x)$ is, of course, local with respect to $u$ for a fixed $x\in E$, then $u(x)-f^{\ast}_{c}(u)$ is local with respect to $u$ when $x$ is fixed. So by (2) of Lemma 3.16, we have that $f(x)=\bigvee\{u(x)-f^{\ast}_{c}(u)~|~u\in H_{cc}(E^{\ast}_c)\}$ (by the fact that $(E, {\mathcal P}_{cc})^\ast_c=H_{cc}(E^\ast_c)$, where $E^\ast_c=(E, {\mathcal P})^\ast_c$) $=\bigvee\{u(x)-f^{\ast}_{c}(u)~|~u\in E^{\ast}_c\}$. \hfill \hfill $\square$ \end{pot} \subsection{Random duality under the locally $L^0$--convex topology with respect to random duality pair} Only the classical duality theory with respect to a duality pair can give a thorough treatment of classical conjugate space theory of locally convex spaces. The theory of random conjugate spaces occupies a central place in the study of $RN$ modules and $RLC$ modules, it is very natural that random duality theory was studied at the previous time in \cite{TXG-dual,TXG-Sur,TXG-XXC}, where many basic results and useful techniques were already obtained. Before 2009, only the $(\varepsilon, \lambda)$--topology was available, so the work in \cite{TXG-dual,TXG-Sur,TXG-XXC} was carried out under this topology, where the family of $L^0$--seminorms plays a key role. In the subsection, we will establish some basic results on random duality theory with respect to the locally $L^0$--convex topology in order to provide an enough framework for the theory of $RLC$ modules and its financial applications, which is motivated from the study of $L^0$--barreled modules. For a sake of convenience, let us introduce the following: \begin{definition} (See \cite{FKV}.) Let $(E, {\mathcal T})$ be a locally $L^0$--convex $L^0({\mathcal F}, K)$--module. An $L^0$--balanced, $L^0$--absorbent, $L^0$--convex and closed subset of $E$ is called an $L^0$--barrel. $(E, {\mathcal T})$ is called an $L^0$--barreled module if every $L^0$--barrel is a neighborhood of $\theta$. \end{definition} Further, D. Filipovi\'{c}, M. Kupper and N. Vogelpoth established the continuity and subdifferentiability theorems for a proper lower semicontinuous $L^0$--convex function defined on an $L^0$--barreled module in \cite{FKV}. Then we naturally ask: what is a characterization for a locally $L^0$--convex $L^0({\mathcal F}, K)$--module to be $L^0$--barreled? In particular, is a ${\mathcal T}_c$--complete $RN$ module $L^0$--barreled under the locally $L^0$--convex topology induced by its $L^0$--norm? Specially, is $L^p_{\mathcal F}(\mathcal E)$ $L^0$--barreled? We at once realized that these problems are rather complicated. Our study leads to the following: \begin{definition} Let $(E, {\mathcal T})$ be a locally $L^0$--convex $L^0({\mathcal F}, K)$--module. $(E,$ ${\mathcal T})$ is called an $L^0$--pre-barreled module if every $L^0$--barrel with the countable concatenation property is a neighborhood of $\theta$. \end{definition} The notion of an $L^0$--pre-barreled module is weaker than that of an $L^0$--barreled module. Although we have not yet provided a perfect characterization for a locally $L^0$--convex $L^0({\mathcal F}, K)$--module to be $L^0$--barreled or $L^0$--pre-barreled, in Subsection 3.3.3 we will prove that every $RLC$ module $(E, {\mathcal P})$ such that $E$ has the countable concatenation property is $L^0$--pre-barreled under the locally $L^0$--convex topology induced by ${\mathcal P}$ iff ${\mathcal T}_c$ is the topology of random uniform convergence on the family of $\sigma_c(E^\ast_c, E)$--bounded sets of $E^\ast_c$. In particular, every ${\mathcal T}_c$--complete $RN$ module $(E, \|\cdot\|)$ is $L^0$--pre-barreled under its locally $L^0$--convex topology if $E$ has the countable concatenation property. Further, in Section 3.4 we will establish the continuity and subdifferentiability theorems for a proper lower semicontinuous $L^0$--convex function defined on an $L^0$--pre-barreled module $(E, {\mathcal T})$ such that $E$ has the countable concatenation property. Thus the notion of an $L^0$--pre-barreled module is more suitable for the study of conditional risk measures. \subsubsection{Random compatible locally $L^0$--convex topology} Let $X$ and $Y$ be two $L^0({\mathcal F}, K)$--modules. A mapping $\langle\cdot, \cdot\rangle\colon X\times Y\to L^0({\mathcal F}, K)$ is said to be $L^0({\mathcal F}, K)$--bilinear function ( briefly, $L^0$--bilinear function if there is not any possible confusion ) if both $\langle x, \cdot\rangle\colon Y\to L^0({\mathcal F}, K)$ and $\langle\cdot, y\rangle\colon X\to L^0({\mathcal F}, K)$ are $L^0$--linear functions for all $x\in X$ and $y\in Y$. \begin{definition} (See \cite{TXG-dual,TXG-Sur,TXG-XXC}.) Two $L^0({\mathcal F}, K)$--modules $X$ and $Y$ are called a random duality pair over $K$ with base $(\Omega, {\mathcal F}, P)$ with respect to the $L^0$--bilinear function $\langle\cdot,\cdot\rangle\colon X\times Y\to L^0({\mathcal F}, K)$ if the following are satisfied:\\ (1). $\langle x, y\rangle=0$ for all $y\in Y$ iff $x=\theta$;\\ (2). $\langle x, y\rangle=0$ for all $x\in X$ iff $y=\theta$. \end{definition} Usually, if $X,~Y$ and $\langle\cdot,\cdot\rangle$ satisfy Definition 3.19, then we simply say that $\langle X, Y\rangle$ is a random duality pair over $K$ with base $(\Omega, {\mathcal F}, P)$. Let $X^\#$ denote the $L^0({\mathcal F}, K)$--module of $L^0$--linear functions from an $L^0({\mathcal F}, K)$--module $X$ to $L^0({\mathcal F}, K)$. It is clear that $X^\#$ has the countable concatenation property. If $\langle X, Y\rangle$ is a random duality pair, we always identify each $x\in X$ with $\langle x, \cdot\rangle\in Y^\#$, namely regard $X$ as a submodule of $Y^\#$, thus for any subset $G\subset X$, we always use $H_{cc}(G)$ for the countable concatenation hull of $G$ in $Y^\#$, which would not cause any possible confusion. \begin{definition} Let $\langle X, Y\rangle$ be a random duality pair over $K$ with base $(\Omega, {\mathcal F}, P)$. A family ${\mathcal P}$ of $L^0$--seminorms on $X$ is called a random compatible family with $Y$ with respect to the locally $L^0$--convex topology ${\mathcal T}_c$ induced by ${\mathcal P}$ if $(X, {\mathcal P})$ becomes an $RLC$ module over $K$ with base $(\Omega, {\mathcal F}, P)$ such that $(E, {\mathcal P})^\ast_c=Y$, in which case we also say that ${\mathcal T}_c$ is a random compatible locally $L^0$--convex topology with $Y$. \end{definition} \begin{remark} In \cite{TXG-XXC}, a family ${\mathcal P}$ of $L^0$--seminorms is called a random compatible family with $Y$ with respect to the $(\varepsilon, \lambda)$--topology ${\mathcal T}_{\varepsilon, \lambda}$ induced by ${\mathcal P}$ if $(X, {\mathcal P})$ becomes an $RLC$ module such that $(E, {\mathcal P})^\ast_{\varepsilon, \lambda}=Y$. \end{remark} Let $\langle X, Y\rangle$ be a random duality pair and $\sigma(X, Y)=\{|\langle\cdot, y\rangle|\colon y\in Y\}$, then $\sigma(X, Y)$ is a family of $L^0$--seminorms on $X$ such that $(X, \sigma(X, Y))$ becomes an $RLC$ module. In the sequel, $\sigma_c(X, Y)$ and $\sigma_{\varepsilon, \lambda}(X, Y)$ always denote the locally $L^0$--convex topology and the $(\varepsilon, \lambda)$--topology induced by $\sigma(X, Y)$, respectively. \begin{theorem} Let $\langle X, Y\rangle$ be a random duality pair over $K$ with base $(\Omega, {\mathcal F}, P)$. Then $\sigma_c(X, Y)$ is a random compatible topology with $Y$. \end{theorem} Proof of Theorem 3.22 needs Lemma 3.23 below. \begin{lemma} $($See \cite{TXG-dual,TXG-XXC}$.)$ Let $E$ be an $L^0({\mathcal F}, K)$--module. If $f_1,f_2,\cdots,f_n$ and $g$ are $n+1$ $L^0$--linear functions from $E$ to $L^0({\mathcal F}, K)$, then there are $\xi_1,\xi_2,\cdots,\xi_n\in L^0({\mathcal F}, K)$ such that $g=\sum^n_{i=1}\xi_if_i$ iff $\bigcap^n_{i=1}N(f_i)\subset N(g)$, where $N(f_i)=\{x\in E~|~f_i(x)=0\}$ ( $1\leq i\leq n$ ) and $N(g)=\{x\in E~|~g(x)=0\}$. \end{lemma} We can now prove Theorem 3.22. \newproof{pot1}{Proof of Theorem 3.22} \begin{pot1} Since it is obvious that $Y\subset (X, \sigma(X, Y))^\ast_c$, it remains to prove that $(X, \sigma(X, Y))^\ast_c \subset Y$. Let $f\in (X, \sigma(X, Y))^\ast_c$, then by Proposition 2.19 there are $\xi\in L^0_+({\mathcal F})$ and $y_1,y_2,\cdots, y_n\in Y$ such that $|f(x)|\leq \xi (\bigvee\{|\langle x, y_i\rangle|\colon 1\leq i\leq n\})$ for all $x\in X$. By Lemma 3.23, there are $\xi_1,\xi_2,\cdots,\xi_n\in L^0({\mathcal F}, K)$ such that $f=\sum^n_{i=1}\xi_i\langle x, y_i\rangle$ for all $x\in X$. Let $y=\sum^n_{i=1}\xi_i y_i$, then $f=y\in Y$. \hfill $\square$ \end{pot1} \begin{remark} In \cite{TXG-XXC}, since we employed the $(\varepsilon, \lambda)$--topology, we proved that $(X,\sigma(X, Y))^\ast_{\varepsilon, \lambda}=H_{cc}(Y)$, which motivates us to find out Theorem 2.30. In \cite{TXG-XXC}, $Y$ is regular with respect to $X$ if, for each sequence $\{y_n, n\in N\}$ and each countable partition $\{A_n, n\in N\}$ of $\Omega $ to ${\mathcal F}$, there is $y\in Y$ such that $\langle x, y\rangle=\sum^{\infty}_{n=1}{\tilde I}_{A_n}\langle x, y_n\rangle$ for all $x\in X$, which implies that ${\tilde I}_{A_n}\langle x, y\rangle={\tilde I}_{A_n}\langle x, y_n\rangle$ for all $x\in X$ and $n\in N$, namely ${\tilde I}_{A_n}y={\tilde I}_{A_n}y_n$ for each $n\in N$, that is to say, $Y$ has the countable concatenation property. Thus, for a random duality pair $\langle X, Y\rangle$, `` $Y$ is regular '' and `` $Y$ has the countable concatenation property '' are the same thing. \end{remark} \begin{theorem} Let $\langle X, Y\rangle$ be a random duality pair. Then there is a strongest one in all the random compatible locally $L^0$--convex topologies with $Y$. \end{theorem} \begin{proof} By Corollary 2.33, one can see that the proof is completely similar to the one of the corresponding classical case, so is omitted. \hfill \hfill $\square$ \end{proof} Let $\langle X, Y\rangle$ be a random duality pair. For a subset $A$ of $X$, $A^0:=\{y\in Y|~|\langle a, y\rangle|\leq 1~\hbox{for all $a\in A$~}\}$ is called the polar of $A$ in $Y$. Similarly, one can define the polar of a subset $B$ of $Y$ in $X$. \begin{theorem} $(${\bf Random bipolar theorem}$)$ Let $\langle X, Y\rangle$ be a random duality pair over $K$ with base $(\Omega, {\mathcal F}, P)$ such that $X$ has the countable concatenation property. Then, for any subset $A$ of $X$, we have that $A^{00}=[H_{cc}(\Gamma (A))]^-_{\mathcal T}$ for each random compatible topology ${\mathcal T}$ with $Y$, where $\Gamma(A)$ denotes the $L^0$--balanced and $L^0$--convex hull of $A$ and $[H_{cc}(\Gamma (A))]^-_{\mathcal T}$ the ${\mathcal T}$--closure of $H_{cc}(\Gamma (A))$. \end{theorem} \begin{proof} Since $(X, {\mathcal T})^\ast_c=Y$, ${\mathcal T}\supset \sigma_c(X, Y)$. On the other hand, it is obvious that $A^{00}$ is an $L^0$--balanced, $L^0$--convex and $\sigma_c(X, Y)$--closed set with the countable concatenation property, so $A^{00}\supset [H_{cc}(\Gamma (A))]^-_{\sigma_c(X, Y)}\supset [H_{cc}(\Gamma (A))]^-_{\mathcal T}$. If there is $x\in A^{00}\setminus [H_{cc}(\Gamma (A))]^-_{\mathcal T}$, then by Corollary 3.9 and Remark 3.11 there is $y\in (X, {\mathcal T})^\ast_c=Y$ such that $|\langle x, y\rangle|\nleqslant 1$ and $\bigvee\{|\langle a, y\rangle|\colon a\in A\}\leq 1$, which is impossible. \hfill \hfill $\square$ \end{proof} \begin{remark} The classical bipolar theorem is an elegant result and hence frequently employed in the study of classical duality theory. However, the random bipolar theorem under the locally $L^0$--convex topology, namely Theorem 3.26 has the complicated form and also requires $X$ to have the countable concatenation property, so we do our best to avoid the use of it except in Subsection 3.3.3 where we are forced to use it to characterize a class of $L^0$--pre-barreled modules. In \cite{TXG-XXC} we proved a random bipolar theorem under the $(\varepsilon, \lambda)$--topology with the same shape as the classical bipolar theorem, but the countable concatenation property of $Y$ is required. To sum up, we are always forced to look for new methods in order to obtain some most refined results on random duality theory. \end{remark} It is time for us to speak of random compatible invariants. Corollary 3.9 shows that any closed $L^0$--convex sets ( in particular, any $L^0$--barrels ) with the countable concatenation property are random compatible invariants with respect to every random duality pair. Theorem 2.35 shows that the same is true for bounded sets in the sense of the locally $L^0$--convex topology. \subsubsection{Random admissible topology} \begin{definition} Let $\langle X, Y\rangle$ be a random duality pair over $K$ with base $(\Omega, {\mathcal F}, P)$ and ${\mathcal A}$ a family of $\sigma_c(Y, X)$--bounded sets of $Y$. For any $ A\in {\mathcal A}$, the $L^0$--seminorm $\|\cdot\|_A\colon X\to L^0_+({\mathcal F})$ is defined by $\|x\|_A=\bigvee\{|\langle x, a\rangle|\colon a\in A\}$ for all $x\in X$. Then the locally $L^0$--convex topology induced by the family $\{\|\cdot\|_A\colon A\in {\mathcal A}\}$ of $L^0$--seminorms, denoted by ${\mathcal T}_{\mathcal A}$, is called the topology of random uniform convergence on ${\mathcal A}$. \end{definition} \begin{definition} Let $\langle X, Y\rangle$ and ${\mathcal A}$ be the same as in Definition 3.28. ${\mathcal T}_{\mathcal A}$ is said to be random admissible if ${\mathcal T}_{\mathcal A}\supset \sigma_c(X, Y)$, in which case ${\mathcal A}$ is said to be random admissible. If $\mathcal{T}_{\mathscr{A}}$ is random compatible, namely $(X,\mathcal{T}_{\mathscr{A}})^{\ast}_c=Y$, then $\mathscr{A}$ is also said to be random compatible. \end{definition} As usual, let us first study $\mathcal{T}_{\mathscr{A}}$. \begin{proposition} Let $\langle X,Y\rangle$ and $\mathscr{A}$ be the same as in Definition 3.29. Then the following are equivalent: \noindent (1). $\mathcal{T}_{\mathscr{A}}$ is Hausdorff. \noindent (2). $\bigcup\mathscr{A}:=\bigcup\{A:A\in\mathscr{A}\}$ is total, namely $\langle x,y\rangle=0$ for all $y\in \bigcup\mathscr{A}$ implies $x=\theta$. \noindent (3). $Span\mathscr{A}$ (the submodule generated by $\bigcup\mathscr{A}$) is $\sigma_{\varepsilon,\lambda}(Y,X)$--dense in $Y$. \noindent (4). $H_{cc}(Span\mathscr{A})$ is $\sigma_{c}(H_{cc}(Y),X)$--dense in $H_{cc}(Y)$. \end{proposition} \begin{proof} (1)$\Leftrightarrow$(2), (3)$\Rightarrow$(2) and (4)$\Rightarrow$(2) are all obvious. (2)$\Rightarrow$(3). By Corollary 3.10 and Remark 3.11, one can complete the proof by the same method as used in the classical case. (3)$\Rightarrow$(4). By applying (2) of Theorem 2.28 to $Span\mathscr{A}$ and $(H_{cc}(Y),\sigma_{c}(H_{cc}(Y),X))$, we have the following relations:$$[H_{cc}(Span\mathscr{A})]^{-}_{\sigma_{c}(H_{cc}(X),Y)}$$$$=[H_{cc}(Span\mathscr{A})]^{-}_{\sigma_{\varepsilon,\lambda}(H_{cc}(X),Y)}$$ $$=[Span\mathscr{A}]^{-}_{\sigma_{\varepsilon,\lambda}(H_{cc}(X),Y)}~~~~~~~$$$$\supset H_{cc}(Y)~~~~~~~~~~~~~~~~~~~~~~~~$$ (by applying (3) to $H_{cc}(Y)$). \hfill $\square$ \end{proof} Although random bipolar theorem does not necessarily hold for all random duality pairs, (2) of Lemma 3.31 below can complement this point. \begin{lemma}\noindent Let $\langle X,Y\rangle$ be a random duality pair. Then we have: \noindent (1). $A\subset Y$ is $\sigma_{c}(Y,X)$--bounded iff $A^{0}$ is a $\sigma_{c}(X,Y)$--$L^0$--barrel. \noindent (2). For any $\sigma_{c}(Y,X)$--bounded set $A\subset Y$, $\|\cdot\|_B=\|\cdot\|_{B^{00}}$ (and hence $B^{00}$ is also $\sigma_{c}(Y,X)$--bounded), where $\|\cdot\|_B$ and $\|\cdot\|_{B^{00}}$ are defined as in Definition 3.27. \end{lemma} \begin{proof} (1) is clear. (2). Since $B^{00}\supset B$, it is obvious that $\|\cdot\|_{B^{00}}\geq\|\cdot\|_B$. Conversely, if $\|x\|_B\leq 1$, then $x\in B^{0}$, and hence $\|x\|_{B^{00}}\leq 1$, which implies $\|\cdot\|_{B^{00}}\leq\|\cdot\|_B$. \hfill\hfill $\square$ \end{proof} \begin{definition} Let $\langle X,Y\rangle$ be a random duality pair over $K$ with base $(\Omega,\mathcal{F},P)$ and $\mathscr{B}$ a family of $\sigma_{c}(Y,X)$--bounded sets of $Y$. $\mathscr{B}$ is saturated if the following are satisfied: \noindent (a). If $A\subset B$ for some $B\in\mathscr{B}$ , then $A\in \mathscr{B}$; \noindent (b). $A,B\in\mathscr{B}\Rightarrow A\bigcup B\in\mathscr{B}$; \noindent (c). $B\in\mathscr{B}\Rightarrow B^{00}\in\mathscr{B}$; \noindent (d). $\lambda B\in\mathscr{B}$ for all $\lambda\in L^0(\mathcal{F},K)$ and $B\in\mathscr{B}$. \end{definition} In the classical definition of a saturated family (which amouts to the case when $\mathcal{F}=\{\Omega,\emptyset\}$), the above (c) in Definition 3.32 is defined as ``$B\in\mathscr{B}\Rightarrow [\Gamma(B)]^{-}_{\sigma(Y,X)}\in\mathscr{B}$". But, generally, we only have the relation that $B^{00}\supset[\Gamma(B)]^{-}_{\sigma_c(Y,X)}$. Although the random bipolar theorem shows that $B^{00}=[H_{cc}(\Gamma(B))]^{-}_{\sigma_{c}(Y,X)}$ if $Y$ has the countable concatenation property, we would like to introduce the notion of a saturated family for an arbitrary random duality pair, so we choose Definition 3.32 to meet all our requirements. Let $\langle X,Y\rangle$ be a random duality pair, in this paper we always denote by $\mathscr{B}(Y,X)$ the family of $\sigma_{c}(Y,X)$--bounded sets of $Y$ and $\beta(X,Y)=\mathcal{T}_{\mathscr{B}(Y,X)}$. By (2) of Lemma 3.31, $\mathscr{B}(Y,X)$ is saturated. It is also obvious that $\beta(X,Y)$ is the strongest random admissible topology. For a family $\mathscr{A}$ of $\sigma_{c}(Y,X)$--bounded sets of $Y$, $\mathscr{A}^{s}$ denotes the saturated hull of $\mathscr{A}$, namely the smallest saturated family containing $\mathscr{A}$. It is easy to see that $\mathscr{A}^{s}=\{B\subset Y~|~$ there are $\lambda_1,\lambda_2,\cdots,\lambda_n\in L^{0}(\mathcal{F},K)$ and $A_1,A_2,\cdots,A_n\in\mathscr{A}$ such that $B\subset(\bigcup_{i=1}^{n}\lambda_i A_i)^{00}\}$, again by (2) of Lemma 3.31 one can easily see that $\mathcal{T}_{\mathscr{A}}=\mathcal{T}_{\mathscr{A}^s}$. \begin{proposition} Let $\langle X,Y\rangle$ be a random duality pair over $K$ with base $(\Omega,\mathcal{F},P)$ and $\mathscr{A}$ and $\mathscr{B}$ two family of $\sigma_c(Y,X)$--bounded sets of $Y$ such that $\mathscr{B}$ is saturated. Then, $\mathcal{T}_{\mathscr{A}}\subset\mathcal{T}_{\mathscr{B}}$ iff $\mathcal{A}\subset\mathcal{B}$. \end{proposition} \begin{proof} If $\mathcal{T}_{\mathscr{A}}\subset\mathcal{T}_{\mathscr{B}}$, then for each $A\in\mathscr{A}$ there are $\xi\in L^{0}_{+}(\mathcal{F})$ and a finite subfamily $\{B_i~|~1\leq i\leq n\}$ of $\mathscr{B}$ such that $\|x\|_A\leq\xi(\bigvee\{\|x\|_{B_i}:1\leq i\leq n\})=\|x\|_B$ for all $x\in X$, where $B=\xi(\bigcup_{i=1}^{n}B_i)\in \mathscr{B}$. Thus $A\subset A^{00}\subset B^{00}\in\mathscr{B}$, which has showed that $A\in\mathscr{B}$. The converse is obvious. \hfill $\square$ \end{proof} \begin{corollary} Let $\langle X,Y\rangle$ be a random duality pair and $\mathscr{A}$ is a saturated family of $\sigma_{c}(Y,X)$--bounded sets of $Y$. Then $\mathcal{T}_{\mathscr{A}}$ is random admissible iff $\bigcup\mathscr{A}=Y$. \end{corollary} \begin{proof} Let $Y_f$ denote the family of finite subsets of $Y$, then $\sigma_{c}(X,Y)=\mathcal{T}_{Y_f}$. So, $\mathcal{T}_{\mathscr{A}}$ is random admissible iff $\mathcal{T}_{Y_f}\subset\mathcal{T}_{\mathscr{A}}$ iff $Y_f\subset \mathscr{A}$ iff $\bigcup \mathscr{A}=Y$. \hfill $\square$ \end{proof} \begin{theorem} Let $\langle X,Y\rangle$ be a random duality pair over $K$ with base $(\Omega,\mathcal{F},P)$. Then a locally $L^{0}$--convex topology $\mathcal{T}$ on $X$ is a topology of random uniform convergence iff $\mathcal{T}$ has a local base $\mathcal{B}$ at $\theta$ such that each $U\in\mathcal{B}$ is a $\sigma_c(X,Y)$--$L^{0}$--barrel. \end{theorem} \begin{proof} If $\mathcal{T}=\mathcal{T}_{\mathscr{A}}$ for some family $\mathscr{A}$ of $\sigma_c(X,Y)-$bounded sets of $Y$, we can suppose that $\mathscr{A}$ is saturated, then $\mathcal{B}=\{A^{0}:A\in\mathscr{A}\}$ is a local base at $\theta$ of $\mathcal{T}$ such that each $A^{0}$ is a $\sigma_{c}(X,Y)$--$L^{0}$--barrel. Conversely, let $\mathcal{T}$ have a local base $\mathcal{B}$ at $\theta$ such that each $U\in \mathcal{B}$ is a $\sigma_c(X,Y)$--$L^0$--barrel. Let $\mathscr{A}=\{U^{0}:U\in\mathscr{B}\}$, then each $U^{0}$ is $\sigma_{c}(Y,X)$--bounded since $(U^{0})^{0}\supset U$ is $L^0-$absorbent. Further, we show that $\mathcal{T}=\mathcal{T}_{\mathscr{A}}$ as follows. Since $\mathcal{T}_{\mathscr{A}}$ is induced by $\{\|\cdot\|_{A}:A\in\mathscr{A}\}$ and , for each $U^{0}\in\mathscr{A}$, $\{x\in X~|~\|x\|_{U^{0}}\leq 1\}=U^{00}\supset U$, which shows that $\|\cdot\|_{U^0}$ is $\mathcal{T}$--continuous, namely $\mathcal{T}_{\mathscr{A}}\subset \mathcal{T}$. On the other hand, for each $U\in\mathcal{B}$, $U\subset U^{00}=(U^{0})^{0}$, namely each element $(U^{0})^{0}$ of a local base at $\theta$ of $\mathcal{T}_{\mathscr{A}}$ is a $\mathcal{T}$--neighborhood of $\theta$, that is to say, $\mathcal{T}_{\mathscr{A}}\subset\mathcal{T}$. \hfill $\square$ \end{proof} In the classical case, by the classical bipolar theorem it can be easily established that $\{A^{0}:A\in\mathscr{B}(Y,X)\}$ as the local base at $\theta$ of $\beta(X,Y)$ is exactly the family of $\sigma(X,Y)$--barrels. However, in the random setting, we do not know if $\{A^{0}:A\in\mathscr{B}(Y,X)\}$ as the local base at $\theta$ of $\beta(X,Y)$ is still the family of $\sigma_{c}(X,Y)$--$L^0$--barrels, we only know that for each $\sigma_{c}(X,Y)$--$L^0$--barrel $U$ there is $A(=U^0)\in\mathscr{B}(Y,X)$ such that $U\subset A^{0}$. So, we remind the reader of the following useful result: \begin{theorem} Let $\langle X,Y\rangle$ be a random duality pair such that $X$ has the countable concatenation property. Then the family of $\sigma_c(X,Y)$--$L^0$--barrels with the countable concatenation property forms a local base at $\theta$ of $\beta(X,Y)$. \end{theorem} \begin{proof} By the countable concatenation property of $X$, it is easy to see that $A^0$ has the countable concatenation property for each $A\in\mathscr{B}(Y,X)$, and hence also a $\sigma_{c}(X,Y)$--$L^0$--barrel with this property. On the other hand, for each $\sigma_{c}(X,Y)$--$L^0$--barrel $U$ with the countable concatenation property, then by Theorem 3.26 we have that $U=U^{00}=(U^{0})^{0}$. Since $U^{0}\in\mathscr{B}(Y,X)$, $U\in\{A^{0}:A\in\mathscr{B}(Y,X)\}$. To sum up, the family of $\sigma_{c}(X,Y)$--$L^{0}$--barrels with the countable concatenation property is exactly the local base $\{A^{0}:A\in\mathscr{B}(Y,X)\}$. \hfill $\square$ \end{proof} Theorem 3.37 below shows that the study of random admissible topology is of universal interest in the theory of $RCL$ modules. \begin{theorem} Let $(X,\mathcal{P})$ be an $RLC$ module over $K$ with base $(\Omega,\mathcal{F},P)$ and $\mathcal{E}$ the family of all the subsets $E$ of $X_{c}^{\ast}$ such that $E$ is equicontinuous from $(X,\mathcal{T}_c)$ to $L^{0}(\mathcal{F},K)$ endowed with the locally $L^{0}$--convex topology induced by $|\cdot|$. Then $\mathcal{T}_{c}=\mathcal{T}_{\mathcal{E}}$, where we consider the natural pairing $\langle X,X_{c}^{\ast}\rangle$, then $\mathcal{T}_{\mathcal{E}}$ is, clearly, a random admissible topology. \end{theorem} \begin{proof} It is clear that $E\in\mathcal{E}$ iff there are $\xi\in L^{0}_{+}(\mathcal{F})$ and a finite subset $\mathcal{Q}$ of $\mathcal{P}$ such that $\|x\|_{E}:=\bigvee\{|f(x)|:f\in E\}\leq \xi(\bigvee\{\|x\|:\|\cdot\|\in\mathcal{Q}\})$ for all $x\in X$, so $\mathcal{T}_{\mathcal{E}}\subset\mathcal{T}_{c}$. Conversely, for each $\|\cdot\|\in\mathcal{P}$, let $E=\{f\in X^{\ast}_{c}~|~|f(x)|\leq \|x\|$ for all $x\in X\}$, then from the random Hahn-Banach theorem of \cite{TXG-JFA} one can easily see that $\|\cdot\|=\|\cdot\|_E$, so $\mathcal{T}_{c}\subset\mathcal{T}_{\mathcal{E}}$. \hfill $\square$ \end{proof} \begin{corollary} Let $\langle X,Y\rangle$ be a random duality pair over $K$ with base $(\Omega,\mathcal{F},P)$. Then every random compatible topology $\mathcal{T}$ on $X$ is random admissible. \end{corollary} \begin{proof} By Definition 3.20, there is a family $\mathcal{P}$ of $L^0$--seminorms on $X$ such that $(X,\mathcal{P})$ becomes an $RLC$ module over $K$ with base $(\Omega,\mathcal{F},P)$ and $(X,\mathcal{P})^{\ast}_{c}=Y$, $\mathcal{T}$ is just induced by $\mathcal{P}$, at which time $\langle X,Y\rangle$ is exactly $\langle X,X^{\ast}_{c}\rangle$ and $\mathcal{T}=\mathcal{T}_{\mathcal{E}}$ by Theorem 3.37. \hfill $\square$ \end{proof} The proof of Theorem 3.39 below (namely the resonance theorem) is omitted since it is the same as that of the classical case. \begin{theorem} Let $(E,\mathcal{P})$ be an $RLC$ module over $K$ with base $(\Omega,\mathcal{F},P)$ and $H\subset E^{\ast}_{c}$. Then we have the following: \noindent $(1)$. If $(E,\mathcal{T}_c)$ is $L^0$--barreled, then $H$ is equicontunuous from $(E,\mathcal{T}_c)$ to $(L^{0}(\mathcal{F},K),\mathcal{T}_c)$ iff $H$ is $\sigma_c(E^{\ast}_{c},E)$--bounded. \noindent $(2)$. If $(E,\mathcal{T}_c)$ is $L^0$--pre-barreled and $E$ has the countable concatenation property, then $H$ is equicontinuous from $(E,\mathcal{T}_c)$ to $(L^{0}(\mathcal{F},K),\mathcal{T}_c)$ iff $H$ is $\sigma_c(E^{\ast}_{c},E)$--bounded. \end{theorem} In the classical case, for a locally convex space $(E,\mathcal{T})$, a subset $H\subset E^{\ast}$ is equicontinuous, then it must be $\sigma(E^{\ast},E)$--relatively compact. However the classical Banach-Alaoglu theorem universally fails to hold in the case of $RN$ modules under the $(\varepsilon,\lambda)$--topology (cf. \cite{TXG-Alao}), the same, of course, occurs for the locally $L^{0}$--convex topology, so we can not generalize the construction of the classical Mackey topology to the random setting. \subsubsection{A characterization for a random locally convex module to be $L^0$--pre-barreled} The main result of this subsection is Theorem 3.40 below. \begin{theorem} Let $(E,\mathcal{P})$ be an $RLC$ module over $K$ with base $(\Omega,\mathcal{F},P)$ such that $E$ has the countable concatenation property. Then $(E,\mathcal{T}_c)$ is $L^0$--pre-barreled iff $\mathcal{T}_c=\beta(E,E^{\ast}_c)$. \end{theorem} \begin{proof} $\mathcal{T}_c$ has a local base at $\theta$ consisting of $\{N_{\theta}(\mathcal{Q},\varepsilon)~|~\mathcal{Q}\in\mathcal{P}_f,\varepsilon\in L^{0}_{++}(\mathcal{F})\}$, where $N_{\theta}(\mathcal{Q},\varepsilon)=\{x\in E~|~\|x\|_{\mathcal{Q}}\leq\varepsilon\}$. It is obvious that every $N_{\theta}(\mathcal{Q},\varepsilon)$ is an $\mathcal{T}_c$--$L^0$--barreled with the countable concatenation property. Thus, if $\mathcal{T}_c$ is $L^0$--pre-barreled, then the family of $\mathcal{T}_c$--$L^0$--barrels with the countable concatenation property forms a local base at $\theta$ of $\mathcal{T}_c$. By Corollary 3.9, $L^0$--barrels with the countable concatenation property are random compatible invariants, namely the family of $\mathcal{T}_c$--$L^0$--barrels with the countable concatenation property coincides with the family of $\sigma_c(E,E_c^{\ast})$--$L^0$--barrels with the countable concatenation property, so $\mathcal{T}_c=\beta(E,E_c^{\ast})$ by Theorem 3.36 if $\mathcal{T}_c$ is $L^0$--pre-barreled. Conversely, if $\mathcal{T}_c=\beta(E,E_c^{\ast})$, then ,since every $\mathcal{T}_c$--$L^0$--barrel with the countable concatenation property is $\sigma_c(E,E_c^{\ast})$--$L^0$--barrel by Corollary 3.9, and hence a $\beta(E,E^{\ast}_c)$--neighborhood of $\theta$ by Theorem 3.36, namely a $\mathcal{T}_c$--neighborhood of $\theta$, that is to say, $(E,\mathcal{T}_c)$ is $L^0$--pre-barreled. ~~~~~~~\hfill $\square$ \end{proof} \begin{corollary} Let $(E,\|\cdot\|)$ be a $\mathcal{T}_c$--complete $RN$ module over $K$ with $(\Omega,\mathcal{F},P)$ such that $E$ has the countable concatenation property. Then $(E,\mathcal{T}_c)$ is $L^0$--pre-barreled. \end{corollary} \begin{proof} We only need to verify that $\mathcal{T}_c=\beta(E,E_c^{\ast})$. First, the locally $L^0$--convex topology $\mathcal{T}_c$ induced by $\|\cdot\|$ is a random compatible topology with respect to the natural random duality pair $\langle E,E_c^{\ast}\rangle$, so $\mathcal{T}_c\subset\beta(E,E_c^{\ast})$ by Corollary 3.38. Conversely, $\beta(E,E_c^{\ast})$ is induced by $\{\|\cdot\|_A:A\in \mathscr{B}(E_c^{\ast},E)\}$, please recall that $\|\cdot\|_A:E\rightarrow L^{0}_{+}(\mathcal{F})$ is given by $\|x\|_A=\bigvee\{|f(x)|:f\in A\}$ for all $x\in E$ and $A\in \mathscr{B}(E_c^{\ast},E)$. Thus we only need to prove that each $\|\cdot\|_A$ is continuous from $(E,\mathcal{T}_c)$ to $(L^0(\mathcal{F},K),\mathcal{T}_c)$. $A\in \mathscr{B}(E_c^{\ast},E)$ means that $\{f(x):f\in A\}$ is $\mathcal{T}_c$--bounded in $(L^0(\mathcal{F},K),\mathcal{T}_c)$ for each $x\in E$, then, by Theorem 2.34 $A$ is $\mathcal{T}_c$--bounded in $E_c^{\ast}$, namely a.s. bounded, and hence there is $\xi_A\in L^0_+(\mathcal{F})$ such that $\|f\|\leq\xi_A$ for all $f\in A$. This shows that $\|x\|_A=\bigvee\{|f(x)|:f\in A\}\leq\bigvee\{\|f\|\cdot\|x\|:f\in A\}\leq(\bigvee\{\|f\|:f\in A\})\|x\|\leq\xi_A\|x\|$ for all $x\in E$, namely $\beta(E,E_c^{\ast})\subset \mathcal{T}_c$. \hfill $\square$ \end{proof} \begin{corollary} For each $p\in[1,+\infty]$, $(L^{p}_{\mathcal{F}}(\mathcal{E}),\mathcal{T}_c)$ is $L^0$--pre-barreled. \end{corollary} \begin{proof} Since $L^{p}_{\mathcal{F}}(\mathcal{E})$ is $\mathcal{T}_c$--complete and has the countable concatenation property, then it immediately follows from Corollary 3.41. \hfill $\square$ \end{proof} \subsection*{3.4. Continuity and subdifferentiability theorems in $L^0$--pre-barreled modules} \quad We will state the results in this subsection under the framework of locally $L^0$--convex modules since the proofs of these results do not necessarily depend on the family of $L^0$--seminorms. Continuity and subdifferentiability theorems in $L^0$--barreled modules were already proved in \cite{FKV}. As shown in \cite{FKV}, the proofs in the random setting are very similar to those in the corresponding classical cases. Thus we omit some details of the proofs in order to save space for some discussions on the relation between the topological structure and stratification structure of a locally $L^0$--convex module. In the subsection, a locally $L^0$--convex module (a topological $L^0$--module) means a locally $L^0$--convex $L^0({\mathcal{F},R})$--module (resp., a topological $L^0({\mathcal{F},R})$--module). To state our main results, let us first recall the following: \begin{definition}(See \cite{FKV}.) Let $(E,\mathcal{T})$ be a locally $L^0$--convex module and $f:E\rightarrow\bar{L}^0(\mathcal{F})$ a proper $L^0$--convex function. $f$ is subdifferentiable at $x\in dom(f)$ if there is $u\in E^{\ast}_{c}$ such that $u(y-x)\leq f(y)-f(x)$ for all $y\in E$, at which time $u$ is called a subgradient of $f$ at $x$. The set of subgradients of $f$ at $x$ is denoted by $\partial f(x)$. \end{definition} We can now state our main results as follows: \begin{theorem} Let $(E,\mathcal{T})$ be a real $L^0$--pre-barreled module such that $E$ has the countable concatenation property. Then a proper lower semicontinuous $L^0$--convex function $f:E\rightarrow \bar{L}^{0}(\mathcal{F})$ is continuous on $Int(dom(f)):=$ the interior of $dom(f)$, namely $f$ is continuous from $(Int(dom(f)),\mathcal{T})$ to $(L^0(\mathcal{F}),\mathcal{T}_c)$. \end{theorem} \begin{theorem} Let $(E,\mathcal{T})$ be a real $L^0$--pre-barreled module such that $E$ has the countable concatenation property. Then, for a proper lower semicontinuous $L^0$--convex function $f:E\rightarrow\bar{L}^{0}(\mathcal{F})$, $\partial f(x)\neq\emptyset$ for all $x\in Int(dom(f))$. \end{theorem} To prove Theorem 3.44, we needs the following known lemmas: \begin{lemma} ($See$ \cite{FKV}.) Let $E$ be a topological $L^0$--module. If in some neighborhood of an element $x_{0}\in E$ a proper $L^0$--convex function $f:E\rightarrow\bar{L}^{0}(\mathcal{F})$ is bounded above by some $\xi_0\in L^{0}(\mathcal{F})$, then $f$ is continuous at $x_{0}$. \end{lemma} \begin{lemma} ($See$ \cite{FKV}.) Let $E$ be a topological $L^0$--module and $f:E\rightarrow\bar{L}^{0}(\mathcal{F})$ a proper $L^0$--convex function. Then the following statements are equivalent: \noindent $(1)$. There is a nonempty open set $O\subset E$ on which $f$ is bounded above by some $\xi_0\in L^{0}(\mathcal{F})$. \noindent $(2)$. $f$ is continuous on $Int(dom(f))$ and $Int(dom(f))\neq\emptyset$. \end{lemma} \begin{lemma} ($See$ \cite{FKV}.) Let $E$ be a topological $L^0$--module and $x\in E$. Then every proper $L^0$--convex function $f:Span_{L^0}(x)\rightarrow \bar{L}^{0}(\mathcal{F})$ is continuous on $Int(dom(f))$, where $Span_{L^0}(x)$ is the $L^0$--module spanned by $x$ and endowed with the relative topology. \end{lemma} We can now prove Theorem 3.44. \newproof{pot2}{Proof of Theorem 3.44} \begin{pot2} Assume that there is $x_{0}\in Int(dom(f))$. By translation, we may assume $x_{0}=0$ and further take $Y_0\in L^{0}(\mathcal{F})$ such that $f(0)<Y_0$ on $\Omega$. Since $f$ is lower semicontinuous, the set $C:=\{x\in E~|~f(x)\leq Y_0\}$ is closed. Further, for all $x\in E$, the net $\{\frac{x}{Y}:Y\in L^{0}_{++}(\mathcal{F})\}$ converge to $\theta$. By Lemma 3.48, the restriction of $f$ to $Span_{L^0}(x)$ is continuous at $\theta$, hence $f(\frac{x}{Y})<Y_0$ on $\Omega$ for large $Y$, which means that $C$ is $L^0$--absorbent. Hence $C\bigcap(-C)$ is an $L^{0}$--barrel. Since $E$ has the countable concatenation property and $f$ has the local property, it is easy to observe that $C\bigcap(-C)$ is an $L^0$--barrel with the countable concatenation property and in turn a neighborhood of $\theta\in E$, so $f$ is continuous on $Int(dom(f))$ by Lemma 3.47. \hfill $\square$ \end{pot2} To prove Theorem 3.45, we need the following three lemmas. Lemma 3.49 below is a slight generalization of Lemma 3.17 of \cite{TXG-JFA}, whereas their proofs are the same, so the proof of Lemma 3.49 is omitted. \begin{lemma} Let $(E,\mathcal{T})$ be a locally $L^0$--convex $L^0(\mathcal{F},K)$--module and $A\in \cal{F}$ with $P(A)>0$. If $G$ and $M$ are an open set and a closed set of $E$, respectively, such that $\tilde{I}_A G+\tilde{I}_{A^{c}} G\subset G$ and $\tilde{I}_A M+\tilde{I}_{A^{c}} M\subset M$, then $\tilde{I}_A G$ is relatively open in $\tilde{I}_A E$ and $\tilde{I}_A M$ is relatively closed in $\tilde{I}_A E$. \end{lemma} From Lemma 3.49 one can see that $(\tilde{I}_{A}E,\mathcal{T}|_{\tilde{I}_{A}E})$ is still a locally $L^0$--convex $L^0(\mathcal{F}_A,K)$--module, where $\mathcal{F}_A=A\bigcap\mathcal{F}:=\{A\bigcap B~|~B\in\mathcal{F}\}$ is the $\sigma$--algebra of $(A,\mathcal{F}_A,P(\cdot|A))$. \begin{lemma} Let $(E,\mathcal{T})$ be a locally $L^0$--convex module, $A\in\mathcal{F}$ with $P(A)>0$ and $f:E\rightarrow \bar{L}^0(\mathcal{F})$ a proper $L^0$--convex function. If, we define $f_A:\tilde{I}_{A}E\rightarrow \tilde{I}_{A}\bar{L}^0(\mathcal{F})$ by $f_A(\tilde{I}_{A}x)=\tilde{I}_{A}f(\tilde{I}_{A}x)$ for all $x\in E$, then we have: \noindent $(1)$. For all $x\in dom(f)$, $\tilde{I}_A(x,f(x))\in \partial(epi(f_A))$, where $\partial(epi(f_A))$ denotes the boundary of $epi(f_A)$ in $(\tilde{I}_{A}E,\mathcal{T}|_{\tilde{I}_{A}E})\times(\tilde{I}_{A}L^0(\mathcal{F}),\mathcal{T}_c|$ $_{\tilde{I}_A L^0{(\mathcal{F})}})$. Here, let us recall that $\mathcal{T}_c$ is the locally $L^0$--convex topology on $L^0(\mathcal{F})$ induced by $|\cdot|$. \noindent $(2)$. For all $x\in dom(f)$, $\tilde{I}_{A}(x,f(x))\not\in \tilde{I}_{A}(Int(epi(f)))$. \end{lemma} \begin{proof} (1). It is easy to see that $(x,f(x))\in\partial(epi(f))$ for all $x\in dom(f)$. By the local property of $f$, $\tilde{I}_{A}(x,f(x))=(\tilde{I}_{A}x, \tilde{I}_{A}f(x))=(\tilde{I}_{A}x,f_A(\tilde{I}_{A}x))$ for all $x\in E$. So, if we consider the corresponding problem in $(\tilde{I}_{A}E,\mathcal{T}|_{\tilde{I}_{A}E})$, then we have that $\tilde{I}_{A}(x,f(x))=(\tilde{I}_{A}x,f_A(\tilde{I}_{A}x))\in\partial(epi(f_A))$ for all $x\in dom(f)$. (2). By the above (1), it is , of course, that $\tilde{I}_{A}(x,f(x))\not\in Int_A(epi(f_A))$, where $Int_A(epi(f_A))$ denotes the interior of $epi(f_A)$ in $(\tilde{I}_{A}E,\mathcal{T}|_{\tilde{I}_{A}E})\times(\tilde{I}_{A}L^0(\mathcal{F}),$ $\mathcal{T}_c|_{\tilde{I}_A L^0{(\mathcal{F})}})$. It is obvious that $epi(f_A)=\tilde{I}_A(epi(f))$. By Lemma 3.49, $\tilde{I}_{A}(Int(epi(f)))$ is an open set in $(\tilde{I}_{A}E,\mathcal{T}|_{\tilde{I}_{A}E})\times(\tilde{I}_{A}L^0(\mathcal{F}),$ $\mathcal{T}_c|_{\tilde{I}_A L^0{(\mathcal{F})}})$, so $Int_A(epi(f_A))=Int_A(\tilde{I}_A epi(f))\supset \tilde{I}_A(Int(epi(f)))$, which implies that $\tilde{I}_A(x,$ $f(x))\not\in \tilde{I}_A(Int$ $(epi(f)))$ for all $x\in dom(f)$. \hfill $\square$ \end{proof} Proof of Lemma 3.51 below is the same as that of Lemma 3.14 of \cite{FKV}, so is omitted. \begin{lemma} Let $(E,\mathcal{T})$ be a locally $L^{0}$--convex module and $f:E\rightarrow\bar{L}^{0}(\mathcal{F})$ a proper lower semicontinuous $L^{0}$--convex function. Then $Int(epi(f))\neq\emptyset$ implies $Int(dom(f))\neq\emptyset$. Furthermore, if, in addition, $(E,\mathcal{T})$ is $L^0$--pre-barreled such that $E$ has the countable concatenation property, then $Int$ $(dom(f))$ $\neq\emptyset$ iff $Int(epi(f))\neq\emptyset$. \end{lemma} We can now prove Theorem 3.45. \newproof{pot3}{Proof of Theorem 3.45} \begin{pot3} By Lemmas 3.50 and 3.51 together with Theorem 3.12, one can complete the proof by the same reasoning as in the proof of Theorem 3.7 of \cite{FKV}. \hfill $\square$ \end{pot3} If the hypothesis ``that $(E,\mathcal{T})$ is $L^0$--pre-barreled module such that $E$ has the countable concatenation property" in Theorems 3.44 and 3.45 is replaced by the one ``that $(E,\mathcal{T})$ is $L^0$--barreled", then Theorems 3.44 and 3.45 change to Proposition 3.5 and Theorem 3.7 of \cite{FKV}, respectively, so we naturally present the following open problem: \noindent {\bf Open problem:} If $E$ has the countable concatenation property, then is $(E,\mathcal{T})$ $L^0$--barreled if $(E,\mathcal{T})$ is $L^0$--pre-barreled? \section{Random convex analysis over random locally convex modules under the $(\varepsilon,\lambda)$--topology} \subsection{Lower semicontinuous $L^0$--convex functions under the $(\varepsilon,\lambda)$--topology} \begin{definition} Let $(E,\mathcal{P})$ be an $RLC$ module over $R$ with base $(\Omega,\mathcal{F},P)$ and $f:E\rightarrow \bar{L}^{0}(\mathcal{F})$ a proper $L^{0}$--convex function. $f$ is $\mathcal{T}_{\varepsilon,\lambda}$--lower semicontinuous if $epi(f)$ is closed in $(E,\mathcal{T}_{\varepsilon,\lambda})\times(L^0(\mathcal{F}),\mathcal{T}_{\varepsilon,\lambda})$. \end{definition} As usual, let $(E,\mathcal{P})$ be an $RLC$ module over $R$ with base $(\Omega,\mathcal{F},P)$ and $f:E\rightarrow L^0(\mathcal{F})$, $f$ is $\mathcal{T}_{\varepsilon,\lambda}$--continuous if $f$ is continuous from $(E,\mathcal{T}_{\varepsilon,\lambda})$ to $(L^0(\mathcal{F}),\mathcal{T}_{\varepsilon,\lambda})$. If $\mathcal{T}_{\varepsilon,\lambda}$ is replaced by $\mathcal{T}_{c}$, then we have the notion of $\mathcal{T}_{c}$--continuity, which is just a special case of the notion of continuity for a function from a locally $L^0$--convex module $(E,\mathcal{T})$ to $(L^0(\mathcal{F}),\mathcal{T}_{c})$. As to why we adopt Definition 4.1 for the $\mathcal{T}_{\varepsilon,\lambda}$--lower semicontinuity of an $L^0$--convex function, we interpret this as follows. If we define the $\mathcal{T}_{\varepsilon,\lambda}$--lower semicontinuity of a proper function $f:(E,\mathcal{P})\rightarrow\bar{L}^0({\mathcal{F}})$ via ``$\{x\in E~|~f(x)\leq r\}$ is $\mathcal{T}_{\varepsilon,\lambda}$--closed for all $r\in L^0(\mathcal{F})$", then this notion is too weak to meet some natural needs of other topics as in \cite{TXG-YJY}. If we define $f$ to be lower semicontinuous via ``$\underline{lim}_\alpha f(x_\alpha):=\bigvee_{\beta\in\Gamma}(\bigwedge_{\alpha\geq\beta}f(x_{\alpha}))\geq f(x)$ for all nets $\{x_{\alpha},\alpha\in\Lambda\}$ in $E$ such that it converges in the $(\varepsilon,\lambda)$--topology to some $x\in E$", the notion is, however, meaningless in the random setting, since we can construct a real $RLC$ module $(E,\mathcal{P})$ and a $\mathcal{T}_{\varepsilon,\lambda}$--continuous $L^0$--convex function $f$ from $E$ to $L^0(\mathcal{F})$, whereas $f$ is not a lower semicontinuous function under this notion. In fact, Definition 4.1 has been proved natural and fruitful, see \cite{TXG-YJY} or this subsection. To connect the $L^0$--convexity and ordinary convexity, we introduce the following terminology: \begin{definition} Let $(E,\mathcal{P})$ be an $RLC$ module over $R$ with base $(\Omega,\mathcal{E},P)$, $\mathcal{F}$ a $\sigma$--subalgebra of $\mathcal{E}$ and $f:E\rightarrow \bar{L}^0(\mathcal{E})$. $f$ is $L^0(\mathcal{F})$--convex if $f(\xi x+(1-\xi)y)\leq \xi f(x)+(1-\xi)f(y)$ for all $x,y\in E$ and $\xi\in L^0_+(\mathcal{F})$ with $0\leq\xi\leq 1$. $f$ is $\mathcal{F}$--local if $\tilde{I}_A f(x)=\tilde{I}_A f(\tilde{I}_A x)$ for all $x\in E$ and $A\in \mathcal{F}$. \end{definition} Let $(\Omega,\mathcal{E},P)$ be a probability space and $\mathcal{F}$ a $\sigma$--subalgebra of $\mathcal{E}$. For any $p\in[1,+\infty]$, $L^p(\mathcal{E}):=L^p(\Omega,\mathcal{E},P)$ denotes the ordinary function space. We can also similarly introduce the $L^0(\mathcal{F})$--convexity and $\mathcal{F}$--locality for a function $f$ from $L^p(\mathcal{E})$ to $\bar{L}^0(\mathcal{E})$. \begin{proposition} Let $(E,\mathcal{P})$ be an $RLC$ module over $R$ with base $(\Omega,\mathcal{E},P)$, $f$ a function from $E$ to $\bar{L}^0(\mathcal{E})$ and $\mathcal{F}$ a $\sigma$--subalgebra of $\mathcal{E}$. Then we have the following statements: \noindent (1). If $f$ is $L^0(\mathcal{F})$--convex, then $f$ is convex (namely, $f$ is a convex function in the ordinary sense) and $\mathcal{F}$--local. \noindent (2). If $f$ is a $\mathcal{T}_{\varepsilon,\lambda}$--continuous function from $E$ to $L^0(\mathcal{E})$, then $f$ is $L^0(\mathcal{F})$--convex iff $f$ is convex and $\mathcal{F}$--local. \end{proposition} \begin{proof} (1). It is clear that $f$ is convex, and the proof of the $\mathcal{F}$--local property of $f$ is similar to the proof of necessity of Lemma 3.14. (2). We only need to prove that $f$ is $L^0(\mathcal{F})$--convex if it is convex and $\mathcal{F}$--local. In fact, for any $\xi\in L_{+}^{0}(\mathcal{F})$ with $0\leq\xi\leq 1$, there is a nondecreasing sequence $\{\xi_n,n\in N\}$ of simple functions in $L^{0}_{+}(\mathcal{F})$ such that $0\leq\xi_n\leq\xi$ for each $n\in N$ and $\{\xi_n,n\in N\}$ converges to $\xi$ with respect to the essentially maximal norm $\|\cdot\|_{\infty}$. Since $f$ is convex and $\mathcal{F}$--local, it is easy to see that $f(\xi_n x+(1-\xi_n)y)\leq \xi_n f(x)+(1-\xi_n)f(y)$ for all $x,y\in E$ and $n\in N$, letting $n\rightarrow\infty$ will yield that $f(\xi x+(1-\xi)y)\leq \xi f(x)+(1-\xi)f(y)$. \hfill $\square$ \end{proof} By the same reasoning, one can prove that for a continuous function $f$ from $(L^p(\mathcal{E}),\|\cdot\|_p)$ to $(L^0(\mathcal{E}),\mathcal{T}_{\varepsilon,\lambda})$ or from $(L^p(\mathcal{E}),\|\cdot\|_p)$ to $(L^r(\mathcal{E}),\|\cdot\|_r)$, $f$ is $L^0(\mathcal{F})$--convex iff $f$ is convex and $\mathcal{F}$--local, where $1\leq r,p\leq+\infty$. The above discussions clarify the relation between $L^0$--convexity and ordinary convexity. Theorem 4.4 below further clarifies the relation between the $\mathcal{T}_{\varepsilon,\lambda}$--lower semicontinuity and $\mathcal{T}_{c}$--lower semicontinuity. \begin{theorem} Let $(E,\mathcal{P})$ be an $RLC$ module over $R$ with base $(\Omega,\mathcal{F},P)$ and $f$ a local proper function from $E$ to $\bar{L}^0(\mathcal{F})$. Then we have the following statements: \noindent $(1)$. If $E$ has the countable concatenation property, then $f$ is $\mathcal{T}_{\varepsilon,\lambda}$--lower semicontinuous iff $epi(f)$ is closed in $(E,\mathcal{T}_c)\times(L^{0}(\mathcal{F}),\mathcal{T}_c)$. \noindent $(2)$. If both $E$ and $\mathcal{P}$ have the countable concatenation property, then $f$ is $\mathcal{T}_{\varepsilon,\lambda}$--lower semicontinuous iff $f$ is $\mathcal{T}_{c}$--lower semicontinuous (namely $\{x\in E~|~f(x)\leq r\}$ is $\mathcal{T}_c$--closed for all $r\in L^0(\mathcal{F})$). \end{theorem} \begin{proof} (1). Since $E$ has the countable concatenation property and $f$ is local, $epi(f)$ has the countable concatenation property. Then, by Proposition 2.27, $epi(f)$ is $\mathcal{T}_{\varepsilon,\lambda}$--closed iff it is $\mathcal{T}_{c}$--closed. (2). If bath $E$ and $\mathcal{P}$ have the countable concatenation property, then , by Proposition 3.15, $f$ is $\mathcal{T}_c$--lower semicontinuous iff $epi(f)$ is $\mathcal{T}_c$--closed, namely closed in $(E,\mathcal{T}_c)\times (L^0(\mathcal{F}),\mathcal{T}_c)$, which together with (1) ends the proof of (2). \hfill\hfill $\square$ \end{proof} \subsection{Fenchel-Moreau dual representation theorem under the $(\varepsilon,\lambda)$--topology} Theorem 4.5 below is the main result of this subsection. \begin{theorem} Let $(E,\mathcal{P})$ be an $RLC$ module over $R$ with base $(\Omega,\mathcal{F},P)$ and $f:E\rightarrow \bar{L}^0(\mathcal{F})$ a proper $\mathcal{T}_{\varepsilon,\lambda}$--lower semicontinuous $L^0$--convex function. The $f^{\ast\ast}_{\varepsilon,\lambda}=f$. Here, $f^{\ast}_{\varepsilon,\lambda}:E^{\ast}_{\varepsilon,\lambda}\rightarrow\bar{L}^{0}(\mathcal{F})$ is defined by $f^{\ast}_{\varepsilon,\lambda}(g)=\bigvee\{g(x)-f(x)~|~x\in E\}$ for all $g\in E^{\ast}_{\varepsilon,\lambda}$, and $f^{\ast\ast}_{\varepsilon,\lambda}:E\rightarrow\bar{L}^0(\mathcal{F})$ by $f^{\ast\ast}_{\varepsilon,\lambda}(x)=\bigvee\{g(x)-f^{\ast}_{\varepsilon,\lambda}(g)~|~g\in E^{\ast}_{\varepsilon,\lambda}\}$ for all $x\in E$. \end{theorem} \begin{proof}\noindent It is complete similar to the first part of proof of Theorem 3.13 only by using Theorem 3.6 in the place of Theorem 3.5, so is omitted. \hfill $\square$ \end{proof} \begin{remark} From Theorem 4.5 we can derive Theorem 3.13. In fact, since $\mathcal{P}_{cc}$ and $\mathcal{P}$ induce the same $(\varepsilon,\lambda)$--topology $\mathcal{T}_{\varepsilon,\lambda}$ on $E$, Theorem 4.5 holds for $(E,\mathcal{P}_{cc})$. We first consider the proof of Theorem 3.13 for $(E,\mathcal{P}_{cc})$: let $\mathcal{T}^{\prime}_{c}$ be the locally $L^{0}$--convex topology induced by $\mathcal{P}_{cc}$ and $\mathcal{T}_c$ still denote the locally $L^0$--convex topology induced by $\mathcal{P}$, then $f$ must be $\mathcal{T}^{\prime}_{c}$--lower semicontinuous since $f$ is $\mathcal{T}_c$--lower semicontinuous, and hence also $\mathcal{T}_{\varepsilon,\lambda}$--lower semicontinuous by (2) of Theorem 4.4 since both $E$ and $\mathcal{P}_{cc}$ have the countable concatenation property. Now, by Theorem 4.5 $f^{\ast\ast}_{\varepsilon,\lambda}=f$, further by $E^{\ast}_{\varepsilon,\lambda}=H_{cc}(E^{\ast}_c)$ and (2) of Lemma 3.16, $f^{\ast\ast}_{c}=f^{\ast\ast}_{\varepsilon,\lambda}$, so $f^{\ast\ast}_{c}=f$. We should also point out that Theorem 4.5 is more convenient in use since it has the same shape as the classical Fenchel-Moreau dual Theorem! \end{remark} In classical convex analysis, people very often need to consider the Fenchel-Moreau dual representation theorem for a not necessarily proper extended real-valued function, where the notion of a closed function is important. Let $(E,\mathcal{T})$ be a locally convex space. $f:E\rightarrow[-\infty,+\infty]$ is closed if either $f\equiv+\infty$, or $f\equiv-\infty$, or $f$ is a proper lower semicontinuous, cf. \cite{ET}. Thus we should also define and study closed functions in the random setting. In fact, D. Filipovi\'{c}, M. Kupper and N. Vogepoth already studied the problem for a special class of $RN$ module $L^{p}_{\mathcal{F}}(\mathcal{E})$ for financial applications. Here, we make use of Theorem 4.5 to give a unified treatment for the problem. Let $(E,\mathcal{P})$ be an $RLC$ module over $R$ with base $(\Omega,\mathcal{F},P)$ and $f:E\rightarrow \bar{L}^{0}(\mathcal{F})$ an local function. Let us first give the following notation: $\mathscr{A}=\{A\in \mathcal{F}~|~$there is $x\in E$ such that $\tilde{I}_A f(x)=\tilde{I}_A(-\infty)\}$; $\mathscr{B}=\{A\in \mathcal{F}~|~\tilde{I}_A f=\tilde{I}_A(+\infty),$ namely $\tilde{I}_A f(x)=\tilde{I}_A(+\infty)$ for all $x\in E\}$; $MI(f)=esssup(\mathscr{A})$; $PI(f)=esssup(\mathscr{B})$; $BP(f)=\Omega\setminus(MI(f)\bigcup PI(f))$; $\mathscr{D}=\{A\subset BP(f)~|~A\in \mathcal{F}$ is such that there are $D\in\mathcal{F}$ with $D\subset A$ and $x\in E$ satisfying $f(x)<+\infty$ on $D\}$. Here, $esssup(\mathcal{H})$ denotes the essential supremum of a subfamily $\mathcal{H}$ of $\mathcal{F}$, cf. \cite{FKV,TXG-JFA}. We can think that $MI(f)$ and $PI(f)$ are disjoint. It is obvious that $\tilde{I}_{PI(f)}f=\tilde{I}_{PI(f)}(+\infty)$ and $f(x)>-\infty$ on $BP(f)$ for all $x\in E$. Since $\mathscr{A}$ and $\mathscr{D}$ are upward directed, one can use the essential supremum theorem to prove Proposition 4.7 below. \begin{proposition} We have the following statements: \noindent $(1)$. There are a countable partition $\{A_n,n\in N\}$ of $MI(f)$ to $\mathcal{F}$ and a sequence $\{y_n,n\in N\}$ in $E$ such that $\tilde{I}_{A_n}f(y_n)=\tilde{I}_{A_n}(-\infty)$ for each $n\in N$. \noindent $(2)$. There are a countable partition $\{D_n,n\in N\}$ of $BP(f)$ to $\mathcal{F}$ and a sequence $\{x_n,n\in N\}$ in $E$ such that $f(x_n)<+\infty$ on $D_n$ for each $n\in N$ (namely, each $\tilde{I}_{D_n}f$ is proper). Further, if, in addition, $P(BP(f))>0$, then each $D_n$ can be chosen such that $P(D_n)>0$. \end{proposition} Let us observe that if $E$ has the countable concatenation property then the local property of $f$ can be used to prove: there are $y\in E$ such that $\tilde{I}_{MI(f)}f(y)=\tilde{I}_{MI(f)}(-\infty)$, and $x\in E$ such that $f(x)<+\infty$ on $BP(f)$, namely $\tilde{I}_{BP(f)}f$ is proper. For each $D\in\mathcal{F}$, let $E^{D}=\tilde{I}_{D}E:=\{\tilde{I}_{D}x~|~x\in E\}$ and $\|\cdot\|^{D}=$ the restriction of $\|\cdot\|$ to $E^{D}$ for each $\|\cdot\|\in\mathcal{P}$. Then $(E^{D},\mathcal{P}^{D})$ can , of course, be regarded as an $RLC$ module over $R$ with base $(D,D\bigcap\mathcal{F},P(\cdot|D))$ if $P(D)>0$, where $\mathcal{P}^{D}=\{\|\cdot\|^{D}~|~\|\cdot\|\in\mathcal{P}\}$. Further, $f_D:E^{D}\rightarrow \tilde{I}_{D}\bar{L}^{0}(\mathcal{F})$ is defined by $f_{D}(\tilde{I}_{D}x)=\tilde{I}_{D}f(\tilde{I}_{D} x)$ for all $x\in E$. We can now introduce the notion of a closed function. We can assume, without loss of generality, that $P(BP(f))>0$ for the function $f$ in discussion. \begin{definition} Let $(E,\mathcal{P})$ be an $RLC$ module over $R$ with base $(\Omega,\mathcal{F},P)$, $f:E\rightarrow \bar{L}^{0}(\mathcal{F})$ a local function. Then $f$ is $\mathcal{T}_{\varepsilon,\lambda}$ (resp., $\mathcal{T}_{c}$)-closed if $\tilde{I}_{MI(f)}f=\tilde{I}_{MI(f)}(-\infty)$ and if $f_A$ is $L^{0}(A\cap\mathcal{F})-$convex and $\mathcal{T}_{\varepsilon,\lambda}$ (resp., $\mathcal{T}_{c}$)-lower semicontinuous for each $A\in\mathcal{F}$ with $A\subset BP(f)$ and $P(A)>0$ such that $f_A$ is proper. \end{definition} \begin{remark} First, $A$ in Definition 4.8 universally exists, for example, let $\{D_n,n\in N\}$ be the same as in (2) of Proposition 4.7, then each $f_{D_{n}}$ is proper. Furthermore, if $f$ is a closed function then $f=\tilde{I}_{PI(f)}(+\infty)+\tilde{I}_{MI(f)}(-\infty)+\sum_{n=1}^{\infty}\tilde{I}_{D_n}f$ with each $\tilde{I}_{D_n}f$ (namely $f_{D_n}$) is proper $L^{0}-$convex lower semicontinuous, so our definition of a closed function is not only very similar to the classical definition of a closed function but also more complicated than the latter. By the way, it is easy to see that a closed function must be $L^{0}-$convex. Secondly, the notion of a $\mathcal{T}_c-$closed function in the sense of Definition 4.8 is more general than that introduced in \cite{FKV-appro}: \cite{FKV-appro} only considered the special case when $E=L^{p}_{\cal F}(\cal E)$, in which case $\tilde{I}_{BP(f)}f$ is proper, whereas $\tilde{I}_{BP(f)}f$ is not necessarily proper in our general case and the study of our general case needs a decomposition of $BP(f)$ as in (2) of Proposition 4.7. Besides, \cite{FKV-appro} employed the strongest notion of a $\mathcal{T}_c-$lower semicontinuous function, whereas we employ the weakest one. \end{remark} \begin{proposition} Let $(E,\mathcal{P})$ be the same as in Definition 4.8, $\{f_\alpha,\alpha\in\Gamma\}$ a family of $\mathcal{T}_{\varepsilon,\lambda}$ (resp., $\mathcal{T}_{c}$)-closed functions from $(E,\mathcal{P})$ to $\bar{L}^{0}(\mathcal{F})$ and define $f=\bigvee\{f_\alpha:\alpha\in \Gamma\}$ by $f(x)=\bigvee\{f_\alpha(x):\alpha\in \Gamma\}$ for all $x\in E$. Then $f$ is still $\mathcal{T}_{\varepsilon,\lambda}$ (resp., $\mathcal{T}_{c}$)-closed. \end{proposition} \begin{proof} It is easy to see that $MI(f)=essinf\{MI(f_{\alpha}),\alpha\in\Gamma\}$, $PI(f)=esssup\{PI(f_{\alpha}),\alpha\in\Gamma\}$ and $\tilde{I}_{MI(f)}f=\tilde{I}_{MI(f)}(-\infty)$. It remains to show that $f_A$ is $L^{0}(A\cap\cal F)-$convex and $\mathcal{T}_{\varepsilon,\lambda}$ (resp. $\mathcal{T}_c$)$-$lower semicontinuous for each $A\in\cal F$ with $A\subset BP(f)$ and $P(A)>0$ such that $f_A$ is proper. We only gives the proof for the $(\varepsilon,\lambda)-$topology since the case for the locally $L^{0}-$convex topology is similar. Since each $f_{\alpha}$ is $\mathcal{T}_{\varepsilon,\lambda}-$closed, each $f_{\alpha}$ is $L^{0}-$convex, then $f$ is $L^{0}-$convex, so $f_A$ is $L^{0}(A\cap\cal F)-$convex. Further, since $epi(f_A)=\cap_{\alpha\in\Gamma}epi((f_{\alpha})_A)$, we only need to check that each $epi((f_{\alpha})_A)$ is $\mathcal{T}_{\varepsilon,\lambda}-$closed in $\tilde{I}_A(E\times L^{0}(\cal F))$. In fact, for any fixed $\alpha\in\Gamma$, $A$ must be a subset of $(PI(f_{\alpha}))^c$ since $A\subset BP(f)$, so $A=(A\cap BP(f_{\alpha}))\cup(A\cap MI(f_{\alpha}))$. According to the fact that $\tilde{I}_{MI(f_{\alpha})}f_{\alpha}=\tilde{I}_{MI(f_{\alpha})}(-\infty), epi((f_{\alpha})_A)=epi((f_{\alpha})_{A\cap BP(f_{\alpha})})+\tilde{I}_{A\cap MI(f_{\alpha})}(E\times L^{0}(\cal F))$. Since $f_A$ is proper, it is obvious that $(f_{\alpha})_{A\cap BP(f_{\alpha})}$ is also proper, which shows that $epi((f_{\alpha})_{A\cap BP(f_{\alpha})})$ is $\mathcal{T}_{\varepsilon,\lambda}-$closed in $\tilde{I}_{A\cap BP(f_{\alpha})}(E\times L^{0}(\cal F))$ since $f_{\alpha}$ is a $\mathcal{T}_{\varepsilon,\lambda}-$closed function. Again by noting the fact that $A\cap BP(f_{\alpha})$ and $A\cap MI(f_{\alpha})$ are disjoint we have that $epi((f_{\alpha})_A)$ is $\mathcal{T}_{\varepsilon,\lambda}-$closed.\hfill $\square$ \end{proof} \begin{definition} Let $(E,\mathcal{P})$ and $f$ be the same as in Definition 4.8. The greatest $\mathcal{T}_{\varepsilon,\lambda}$ (resp., $\mathcal{T}_{c}$)-closed function majorized by $f$, denoted by $Cl_{\varepsilon,\lambda}(f)$ (resp., $Cl_c(f)$), is the $\mathcal{T}_{\varepsilon,\lambda}$ (resp., $\mathcal{T}_c$)-closure of $f$. \end{definition} \begin{lemma} Let $(E,\mathcal{P})$ and $f$ be the same as in Definition 4.8. If $f$ is $\mathcal{T}_{\varepsilon,\lambda}$--closed, then $f^{\ast\ast}_{\varepsilon,\lambda}=f$. \end{lemma} \begin{proof} Since $f$ is $\mathcal{T}_{\varepsilon,\lambda}$--closed, it is obvious that $\tilde{I}_{MI(f)}f^{\ast\ast}_{\varepsilon,\lambda}=\tilde{I}_{MI(f)}f=\tilde{I}_{MI(f)}(-\infty)$ and $\tilde{I}_{PI(f)}f^{\ast\ast}_{\varepsilon,\lambda}=\tilde{I}_{PI(f)}f=\tilde{I}_{PI(f)}(+\infty)$. Let $\{D_n,n\in N\}$ be the same as in (2) of Proposition 4.7 with $P(D_n)>0$ for all $n\in N$,then each $f_{D_n}$ is a proper $L^{0}(D_n\bigcap\mathcal{F})$--convex $\mathcal{T}_{\varepsilon,\lambda}$--lower semicontinuous on $E^{D_n}$. It is also obvious that $\tilde{I}_{D_n}f^{\ast\ast}_{\varepsilon,\lambda}=f^{\ast\ast}_{D_n}=f_{D_n}=\tilde{I}_{D_n} f$ for each $n\in N$ by Theorem 4.5, so $f^{\ast\ast}_{\varepsilon,\lambda}=f$. \hfill $\square$ \end{proof} \begin{theorem} Let $(E,\mathcal{P})$ and $f$ be the same as in Definition 4.8. Then $f^{\ast\ast}_{\varepsilon,\lambda}=Cl_{\varepsilon,\lambda}(f)$. \end{theorem} \begin{proof} It is obvious that $f^{\ast\ast}_{\varepsilon,\lambda}\leq f$ and $f^{\ast\ast}_{\varepsilon,\lambda}$ is $\mathcal{T}_{\varepsilon,\lambda}$--closed, so $f^{\ast\ast}_{\varepsilon,\lambda}\leq Cl_{\varepsilon,\lambda}(f)$. On the other hand, $Cl_{\varepsilon,\lambda}(f)\leq f$, then $Cl_{\varepsilon,\lambda}(f)=(Cl_{\varepsilon,\lambda}(f))^{\ast\ast}_{\varepsilon,\lambda}\leq f^{\ast\ast}_{\varepsilon,\lambda}$ by Lemma 4.12. \hfill $\square$ \end{proof} \begin{corollary} Let $(E,\mathcal{P})$ be an $RLC$ module over $R$ with base $(\Omega,\mathcal{F},P)$ such that $E$ has the countable concatenation property and $f:E\rightarrow\bar{L}^{0}(\mathcal{F})$ a local function, then $f^{\ast\ast}_{c}=Cl_{c}(f)$. \end{corollary} \begin{proof} It is similar to Remark 4.6, so is omitted. \hfill $\square$\end{proof} Let $(E,\mathcal{P})$ be an $RLC$ module over $R$ with base $(\Omega,\mathcal{F},P)$ such that $E$ has the countable concatenation property and $f:E\rightarrow\bar{L}^{0}(\mathcal{F})$ a local function, then by (2) of Lemma 3.16 one can easily observe that $f^{\ast\ast}_{\varepsilon,\lambda}= f_c^{\ast\ast}$, furthermore if ,in addition, $\mathcal{P}$ has the countable concatenation property, then $E^{\ast}_{\varepsilon,\lambda}=E^{\ast}_c$, so $f^{\ast}_{\varepsilon,\lambda}=f_c^{\ast}$, in which case we always denote $f^{\ast}_{\varepsilon,\lambda}$ or $f_c^{\ast}$ by $f^{\ast}$. Finally, let us conclude the subsection with the two representation theorems for $L^{p}_{\mathcal{F}}(\mathcal{E})$--condition risk measures, where $L^{p}_{\mathcal{F}}(\mathcal{E})$ are the $RN$ modules as constructed in Example 2.8 ($1\leq p\leq +\infty$). Definition 4.15 below was introduced for the case when $1\leq p<+\infty$, here we also introduce it for the case when $p=+\infty$ since it will be used in this paper. \begin{definition} Let $1\leq p\leq+\infty$. A proper function $f:L^{p}_{\mathcal{F}}(\mathcal{E})\rightarrow \bar{L}^{0}(\mathcal{F})$ is said to be: \noindent (1). monotone if $f(x)\leq f(y)$ for all $x,y\in L^{p}_{\mathcal{F}}(\mathcal{E})$ such that $x\geq y$; \noindent (2). cash invariant if $f(x+y)=f(x)-y$ for all $x\in L^{p}_{\mathcal{F}}(\mathcal{E})$ and $y\in L^{0}(\mathcal{F})$. \noindent Further, a proper, monotone and cash invariant function $f$ from $L^{p}_{\mathcal{F}}(\mathcal{E})$ to $\bar{L}^{0}(\mathcal{F})$ is called an $L^{p}_{\mathcal{F}}(\mathcal{E})$--conditional risk measure. \end{definition} Since $L^{p}_{\mathcal{F}}(\mathcal{E})$ has the countable concatenation property, and when $1\leq p<+\infty$ $(L^{p}_{\mathcal{F}}(\mathcal{E}))^{\ast}\cong L^{q}_{\mathcal{F}}(\mathcal{E})$, where $p$ and $q$ are a pair of H\"{o}lder conjugate numbers. Further, since $(L^{p}_{\mathcal{F}}(\mathcal{E}),|||\cdot|||_p)$ is an $RN$ module, a proper $L^0(\mathcal{F})$--convex function $f$ from $L^{p}_{\mathcal{F}}(\mathcal{E})$ to $\bar{L}^{0}(\mathcal{F})$ is $\mathcal{T}_{\varepsilon,\lambda}$--lower semicontinuous iff it is $\mathcal{T}_{c}$--lower semicontinuous ($1\leq p<+\infty$). In \cite{FKV-appro}, D. Filipovi\'{c}, M. Kupper and N. Vogelpoth essentially used Theorem 3.13 as well as the typical techniques from conditional risk measures to obtain the following representation theorem: \begin{theorem}($See$ \cite{FKV-appro}.) Let $1\leq p<+\infty$ and $1<q\leq +\infty$ be a pair of H\"{o}lder conjugate numbers and $f:L^{p}_{\mathcal{F}}(\mathcal{E})\rightarrow \bar{L}^{0}(\mathcal{F})$ a $\mathcal{T}_c$ (or equivalently, $\mathcal{T}_{\varepsilon,\lambda}$)-lower semicontinuous $L^{0}(\mathcal{F})$--convex $L^{p}_{\mathcal{F}}(\mathcal{E})$--conditional risk measure. Then $f(x)=\bigvee\{E(xy|\mathcal{F})-f^{\ast}(y)~|~y\in L^{q}_{\mathcal{F}}(\mathcal{E})$, $y\leq 0$ and $E(y|\mathcal{F})=-1\}$ for all $x\in L^{p}_{\mathcal{F}}(\mathcal{E})$. \end{theorem} We consider the natural random duality pair $\langle L^{1}_{\mathcal{F}}(\mathcal{E}),L^{\infty}_{\mathcal{F}}(\mathcal{E})\rangle$, since $L^{1}_{\mathcal{F}}(\mathcal{E})$ and $L^{\infty}_{\mathcal{F}}(\mathcal{E})$ both have the countable concatenation property, a proper $L^{0}(\mathcal{F})$--convex function $f$ from $L^{\infty}_{\mathcal{F}}(\mathcal{E})$ to $\bar{L}^{0}(\mathcal{F})$ is $\sigma_{\varepsilon,\lambda}(L^{\infty}_{\mathcal{F}}(\mathcal{E}),L^{1}_{\mathcal{F}}(\mathcal{E}))$--lower semicontinuous iff it is $\sigma_{c}(L^{\infty}_{\mathcal{F}}(\mathcal{E}),L^{1}_{\mathcal{F}}(\mathcal{E}))$--lower semicontinuous. Further, $(L^{\infty}_{\mathcal{F}}(\mathcal{E}), $ $\sigma(L^{\infty}_{\mathcal{F}}(\mathcal{E}),L^{1}_{\mathcal{F}}(\mathcal{E})))^{\ast}_{\varepsilon,\lambda}= (L^{\infty}_{\mathcal{F}}(\mathcal{E}),\sigma(L^{\infty}_{\mathcal{F}}(\mathcal{E}),L^{1}_{\mathcal{F}}(\mathcal{E})))^{\ast}_{c}= L^{1}_{\mathcal{F}}(\mathcal{E})$. Thus by Theorem 3.13 (or Theorem 4.5) and the similar techniques in \cite{FKV-appro}, we can obtain Theorem 4.17 below. \begin{theorem} Let $f:L^{\infty}_{\mathcal{F}}(\mathcal{E})\rightarrow \bar{L}^{0}(\mathcal{F})$ be a $\sigma_{\varepsilon,\lambda}(L^{\infty}_{\mathcal{F}}(\mathcal{E}),L^{1}_{\mathcal{F}}(\mathcal{E}))$ (or equivalently, $\sigma_{c}(L^{\infty}_{\mathcal{F}}(\mathcal{E}),L^{1}_{\mathcal{F}}(\mathcal{E}))$)-lower semicontinuous $L^0(\mathcal{F})$--convex $L^{\infty}_{\mathcal{F}}(\mathcal{E})$-- conditional risk measure. Then $f(x)=\bigvee\{E(xy|\mathcal{F})-f^{\ast}(y)~|~ y\in L^{1}_{\mathcal{F}}(\mathcal{E}), y\leq0$ and $E(y|\mathcal{F})=-1\}$ for all $x\in L^{\infty}_{\mathcal{F}}(\mathcal{E})$. \end{theorem} \subsection{Extension theorem for $L^{\infty}$--conditional risk measures} In the sequel of this paper $(\Omega,\mathcal{E},P)$ always denotes a probability space, $\mathcal{F}$ a $\sigma$--subalgebra of $\mathcal{E}$, $L^{\infty}(\mathcal{E}):=L^{\infty}(\Omega,\mathcal{E},P)$ and $L^{\infty}(\mathcal{F}):=L^{\infty}(\Omega,\mathcal{F},P)$. \begin{definition} (See \cite{Nadal,DS}.) A function $f:L^{\infty}(\mathcal{E})\rightarrow L^{\infty}(\mathcal{F})$ is said to be: \noindent (1). monotone if $f(x)\leq f(y)$ for all $x,y\in L^{\infty}(\mathcal{E})$ such that $x\geq y$; \noindent (2). cash invariant if $f(x+y)=f(x)-y$ for all $x\in L^{\infty}(\mathcal{E})$ and $y\in L^{\infty}(\mathcal{F})$. \noindent Furthermore, a monotone and cash invariant function from $L^{\infty}(\mathcal{E})$ to $L^{\infty}(\mathcal{F})$ is called an $L^{\infty}$--conditional risk measure. \end{definition} Let $\mathcal{P}$ be the set of all the probability measures $Q$ on $(\Omega,\mathcal{E})$ such that $Q$ is absolutely continuous with respect to $P$ on $\mathcal{E}$ and $\mathcal{P}_{\mathcal{F}}=\{Q\in \mathcal{P}~|~Q=P$ on $\mathcal{F}\}$. For an $L^{\infty}$--conditional risk measure $f:L^{\infty}(\mathcal{E})\rightarrow L^{\infty}(\mathcal{F})$, $\alpha:\mathcal{P}_{\mathcal{F}}\rightarrow \bar{L}^{0}(\mathcal{F})$ defined by $\alpha(Q)=\bigvee\{E_{Q}(-x|\mathcal{F})-f(x):x\in L^{\infty}(\mathcal{E})\}$ for all $Q\in \mathcal{P}_{\mathcal{F}}$, is called the random penalty function of $f$, where $E_{Q}(\cdot|\mathcal{F})$ denotes the conditional expectation given $\mathcal{F}$ under $Q$. The following representation theorem is known. \begin{theorem} ($See$ \cite{Nadal,DS}.) Let $f:L^{\infty}(\mathcal{E})\rightarrow L^{\infty}(\mathcal{F})$ be an $L^{0}(\mathcal{F})$--convex $L^{\infty}$--conditional risk measure. Then the following statements are equivalent: \noindent (1). $f$ is continuous from above, namely $f(x_n)\nearrow f(x)$ whenever $x_n\searrow x$; \noindent (2). $f$ has the ``Fatou property": for any bounded sequence $\{x_n,n\in N\}$ which converges P-a.s. to some $x$, then $f(x)\leq \underline{lim}_{n}f(x_n)$; \noindent (3). $f(x)=\bigvee\{E_{Q}(-x|\mathcal{F})-\alpha(Q)~|~Q\in\mathcal{P}_{\mathcal{F}}\}$ for all $x\in L^{\infty}(\mathcal{E})$. \end{theorem} For an $L^{0}(\mathcal(F))$--convex $L^{\infty}$--conditional risk measure $f:L^{\infty}(\mathcal{E})\rightarrow L^{\infty}(\mathcal{F})$, $f^{\ast}:L^{1}_{\mathcal{F}}(\mathcal{E})\rightarrow \bar{L}^{0}(\mathcal{F})$ defined by $f^{\ast}(y)=\bigvee\{E(xy|\mathcal{F})-f(x)~|~x\in L^{\infty}(\mathcal{E})\}$ for all $y\in L^{1}_{\mathcal{F}}(\mathcal{E})$, is called the random conjugate function of $f$, where $E(\cdot|\mathcal{F})$ denotes the conditional expectation given $\mathcal{F}$ under $P$. By identifying $\mathcal{P}_{\mathcal{F}}$ with $\{y~|~y\in L^{1}_{+}(\mathcal{E}),E(y|\mathcal{F})=1\}$, then (3) of Theorem 4.19 amounts to the following (4). (4). $f(x)=\bigvee\{E(xy|\mathcal{F})-f^{\ast}(y)~|~y\in L^{1}(\mathcal{E}),y\leq 0$ and $E(y|\mathcal{F})=-1\}$. \begin{theorem} (\cite{TXG-reflesive,TXG-SBL}.) Let $(E,\|\cdot\|)$ be an $RN$ module over $K$ with base $(\Omega,\mathcal{F},P)$ and $1\leq p\leq +\infty$. Let $L^{p}(E)=\{x\in E~|~\|x\|_{p}<+\infty\}$, where $\|\cdot\|_{p}:E\rightarrow[0,+\infty]$ is defined by \[ \|x\|_{p}=\left\{ \begin{array}{ll} (\int_{\Omega} \|x\|^{p}dP)^{\frac{1}{p}}, &\mbox{when $1\leq p<+\infty$};\\ inf\{M\in[0,+\infty]~|~\|x\|\leq M\}, &\mbox{when $p=+\infty$,} \end{array} \right. \]\\ for all $x\in E$. \noindent Then $(L^{p}(E),\|\cdot\|_{p})$ is a normed space and $L^{p}(E)$ is $\mathcal{T}_{\varepsilon,\lambda}$--dense in $E$. \end{theorem} \begin{theorem} Let $f:L^{\infty}(\mathcal{E})\rightarrow L^{\infty}(\mathcal{F})$ be an $L^{\infty}$--conditional risk measure. Then there is a unique $L^{\infty}_{\mathcal{F}}(\mathcal{E})$--conditional risk measure $\bar{f}:L^{\infty}_{\mathcal{F}}(\mathcal{E})\rightarrow L^0(\cal{F})$ such that $|\bar{f}(x)-\bar{f}(y)|\leq |||x-y|||_{\infty}$ for all $x,y\in L^{\infty}_{\mathcal{F}}(\mathcal{E})$ and $\bar{f}|_{L^{\infty}(\mathcal{E})}=f$. \end{theorem} \begin{proof} Let us first recall the $L^{0}$--norm $|||\cdot|||_{\infty}:L^{\infty}_{\mathcal{F}}(\mathcal{E})\rightarrow L^0_+(\cal{F})$, which is defined by $|||x|||_{\infty}=\bigwedge\{\xi\in \bar{L}^{0}_{+}(\mathcal{F})~|~|x|\leq\xi\}$ for all $x\in L^{\infty}_{\mathcal{F}}(\mathcal{E})$, then it is obvious that $|||x|||_{\infty}\in L^{\infty}_{+}(\mathcal{F})$ for all $x\in L^{\infty}(\mathcal{E})$. Since $x=y+x-y\leq y+|x-y|\leq y+|||x-y|||_{\infty}$ for all $x,y\in L^{\infty}(\mathcal{E})$, $f(x)\geq f(y+|||x-y|||_{\infty})=f(y)-|||x-y|||_{\infty}$, namely $f(y)-f(x)\leq|||x-y|||_{\infty}$, which shows that $|f(y)-f(x)|\leq|||x-y|||_{\infty}$ for all $x,y\in L^{\infty}(\mathcal{E})$. Since $L^{\infty}(\mathcal{E})=L^{\infty}(L^{\infty}_{\mathcal{F}}(\mathcal{E}))$, $L^{\infty}(\mathcal{E})$ is $\mathcal{T}_{\varepsilon,\lambda}$--dense in $L^{\infty}_{\mathcal{F}}(\mathcal{E})$ by Theorem 4.20. Further, it is clear that $f$ is uniformly $\mathcal{T}_{\varepsilon,\lambda}$--continuous from $(L^{\infty}(\mathcal{E}),|||\cdot|||_{\infty})$ to $(L^{\infty}(\mathcal{F}),|\cdot|)$ (namely $L^{\infty}(\mathcal{E})$ is regarded as a subspace of $(L^{\infty}_{\mathcal{F}}(\mathcal{E}),|||\cdot|||_{\infty})$ and $L^{\infty}({\mathcal{F}})$ as a subspace of $(L^{0}(\mathcal{F}),|\cdot|)$). Thus $f$ has a unique extension $\bar{f}:L^{\infty}_{\mathcal{F}}(\mathcal{E})\rightarrow L^{0}(\mathcal{F})$ such that $|\bar{f}(x)-\bar{f}(y)|\leq|||x-y|||_{\infty}$ for all $x,y\in L^{\infty}_{\mathcal{F}}(\mathcal{E})$. Since $f$ has the $\mathcal{F}$--local property, $\bar{f}$ must have this property. Further, by Lemma 4.22 below every $x\in L^{\infty}_{\mathcal{F}}(\mathcal{E})$ can be expressed as $x=\Sigma_{n=1}^{\infty}\tilde{I}_{A_n}x_n$ for some countable partition $\{A_n,n\in N\}$ of $\Omega$ to $\mathcal{F}$ and some sequence $\{x_n,n\in N\}$ in $L^{\infty}(\mathcal{E})$ and every $y\in L^{0}(\mathcal{F})$ as $y=\Sigma_{n=1}^{\infty}\tilde{I}_{B_n}y_n$ for some countable partition $\{B_n,n\in N\}$ of $\Omega$ to $\cal{F}$ and some sequence $\{y_n,n\in N\}$ in $L^{\infty}(\cal{F})$, where the first series converges in $\mathcal{T}_{\varepsilon,\lambda}$ in $(L^{\infty}(\mathcal{E}),|||\cdot|||_{\infty})$ and the second does in $(L^{0}(\mathcal{F}),|\cdot|)$. Thus one can easily see that $\bar{f}$ is monotone and cash invariant in the sense of Definition 4.15. \hfill $\square$ \end{proof} In Theorem 4.21, when $f$ is $L^{0}(\mathcal{F})$--convex, it is clear that $\bar{f}$ is also $L^{0}(\mathcal{F})$--convex. Lemma 4.22 below is a stronger proposition than the fact that $L^p(\mathcal{E})$ is $\mathcal{T}_{\varepsilon,\lambda}$--dense in $(L^{p}_{\mathcal{F}}(\mathcal{E}),|||\cdot|||_p)$, which will be used in the proofs of Theorems 4.25 and 4.31 below. \begin{lemma} Let $1\leq p\leq+\infty$, then $L^{p}_{\mathcal{F}}(\mathcal{E})=H_{cc}(L^{p}(\mathcal{E}))$.\end{lemma} \begin{proof} Since $L^{p}_{\mathcal{F}}(\mathcal{E})$ has the countable concatenation property, $H_{cc}(L^{p}(\mathcal{E}))\subset L^{p}_{\mathcal{F}}(\mathcal{E})$ is obvious. Conversely, since $L^{p}_{\mathcal{F}}(\mathcal{E})=L^{0}(\mathcal{F})\cdot L^{p}(\mathcal{E}):=\{\xi g~|~\xi\in L^{0}(\mathcal{F})$ and $g\in L^{p}(\mathcal{E})\}$, for any $x=\xi g\in L^{p}_{\mathcal{F}}(\mathcal{E})$ for some $\xi\in L^{0}(\mathcal{F})$ and $g\in L^{p}(\mathcal{E})$, let $\xi^{0}$ be any chosen representative of $\xi$ and $A_n=\{\omega\in\Omega~|~n-1\leq|\xi^{0}(\omega)|<n\}$ for each $n\in N$, then it is clear that $\xi=\Sigma_{n=1}^{\infty}\tilde{I}_{A_n}\xi$, and hence $x=\xi g=(\Sigma_{n=1}^{\infty}\tilde{I}_{A_n}\xi)g=\Sigma_{n=1}^{\infty}\tilde{I}_{A_n}(\tilde{I}_{A_n}\xi g)\in H_{cc}(L^{p}(\mathcal{E}))$ by an easy observation that each $\tilde{I}_{A_n}\xi g\in L^{p}(\mathcal{E})$. \hfill $\square$ \end{proof} \begin{remark} Let $x\in L^{p}_{\mathcal{F}}(\mathcal{E})$, $\{A_n,n\in N\}$ be a countable partition of $\Omega$ to $\mathcal{F}$ and $\{x_n,n\in N\}$ a sequence in $L^p(\mathcal{E})$ such that $x=\Sigma_{n=1}^{\infty}\tilde{I}_{A_n}x_n$. Since it is obvious that $\Sigma_{n=1}^{\infty}\tilde{I}_{A_n}|||x_n|||_p$ converges in probability measure $P$, $\Sigma_{n=1}^{\infty}\tilde{I}_{A_n}x_n$ both converges in $\mathcal{T}_{\varepsilon,\lambda}$ to $x$ and unconditionally converges in $\mathcal{T}_{\varepsilon,\lambda}$ to $x$ in $(L^{p}_{\mathcal{F}}(\mathcal{E}),|||\cdot|||_p)$. \end{remark} Lemma 4.24 below is also crucial for the proofs of Theorems 4.25 and 4.31 below. \begin{lemma} Let $1\leq q \leq+\infty$ and $\gamma$ be any positive number. Then we have the following statements: \noindent (1). Let $G_1=\{y\in L^{q}_{\mathcal{F}}(\mathcal{E})~|~y\leq 0$ and $E(y|\mathcal{F})=-1\}$ and $G_2=\{y\in L^{q}(\mathcal{E})~|~y\leq 0$ and $E(y|\mathcal{F})=-1\}$, then $G_1=H_{cc}(G_2)$. \noindent (2). Let $G_1$ be the same as in (1) above, $G_3=\{y\in L^{q}(\mathcal{E})~|~y\leq 0, E(y|\mathcal{F})=-1$ and $E(|y|^q|\mathcal{F})\in L^{\infty}(\mathcal{F})\}$ and $G_4=\{y\in L^{q}(\mathcal{E})~|~y\leq 0, E(y|\mathcal{F})=-1$ and $E(|y|^q|\mathcal{F})\in L^{\gamma}(\mathcal{F})\}$, then $G_1=H_{cc}(G_3)=H_{cc}(G_4)$. \end{lemma} \begin{proof} (1). It is obvious that $G_1$ has the countable concatenation property and $G_2\subset G_1$, so $H_{cc}(G_2)\subset G_1$. Conversely, let $y$ be a given element in $G_1$, $\xi^0$ any chosen representative of $|||y|||_q$, $A_n=\{\omega\in\Omega~|~n-1\leq\xi^0(\omega)<n\}$ and $y_n=\tilde{I}_{A_n}y+\tilde{I}_{A_n^c}(-1)$ for all $n\in N$, then it is clear that each $y_n\in G_2$ and $y=(\Sigma_{n=1}^{\infty}\tilde{I}_{A_n})y=\Sigma_{n=1}^{\infty}\tilde{I}_{A_n}y=\Sigma_{n=1}^{\infty}\tilde{I}_{A_n}y_n$, so $y\in H_{cc}(G_2)$. Thus $G_1=H_{cc}(G_2)$. (2). It is obvious that $G_3\subset G_4\subset G_1$, so $H_{cc}(G_3)\subset H_{cc}(G_4)\subset G_1$, it remains to prove that $G_1\subset H_{cc}(G_3)$ as follows: let $y\in G_1$, $\xi^0$, $A_n$ and $y_n$ be the same as in the proof of (1) for each $n\in N$, then it is very easy to observe that each $y_n$ also belongs to $G_3$, so $y=\Sigma_{n=1}^{\infty}\tilde{I}_{A_n}y_n\in H_{cc}(G_3)$, which shows that $G_1\subset H_{cc}(G_3)$. \hfill $\square$ \end{proof} \begin{theorem} Let $f:L^{\infty}(\mathcal{E})\rightarrow L^{\infty}(\mathcal{F})$ be an $L^{0}(\mathcal{F})$--convex $L^{\infty}$--conditional risk measure and $\bar{f}:L^{\infty}_{\mathcal{F}}(\mathcal{E})\rightarrow L^{0}(\mathcal{F})$ the unique extension as determined in Theorem 4.21. Then the following statements are equivalent: \noindent $(1)$. $f$ is continuous from above; \noindent $(2)$. $f$ has the Fatou property; \noindent $(3)$. $f(x)=\bigvee\{E(xy|\mathcal{F})-f^{\ast}(y)~|~y\in L^{1}(\mathcal{E}), y\geq 0$ and $E(y|\mathcal{F})=-1\}$ for all $x\in L^{\infty}(\mathcal{E})$; \noindent $(4)$. $f(x)=\bigvee\{E(xy|\mathcal{F})-f^{\ast}(y)~|~y\in L^{1}_{\mathcal{F}}(\mathcal{E}),y\leq 0$ and $E(y|\mathcal{F})=-1\}$ for all $x\in L^{\infty}(\mathcal{E})$; \noindent $(5)$. $\bar{f}(x)=\bigvee\{E(xy|\mathcal{F})-\bar{f}^{\ast}(y)~|~y\in L^{1}_{\mathcal{F}}(\mathcal{E}),y\leq 0$ and $E(y|\mathcal{F})=-1\}$ for all $x\in L^{\infty}_{\mathcal{F}}(\mathcal{E})$; \noindent $(6)$. $\bar{f}$ is $\sigma_{\varepsilon,\lambda}(L^{\infty}_{\mathcal{F}}(\mathcal{E}),L^{1}_{\mathcal{F}}(\mathcal{E}))$ (or equivalently, $\sigma_{c}(L^{\infty}_{\mathcal{F}}(\mathcal{E}),$ $L^{1}_{\mathcal{F}}(\mathcal{E}))$)-lower semicontinuous. \end{theorem} \begin{proof} (1)$\Leftrightarrow$(2)$\Leftrightarrow$(3) is just Theorem 4.19. (3)$\Leftrightarrow$(4). The equivalence relation is easily seen by applying (1) of Lemma 4.24 for $q=1$ and (2) of Lemma 3.16 since $E(xy|\mathcal{F})-f^{\ast}(y)$ is a local function of $y$ for each fixed $x\in L^{\infty}(\mathcal{E})$. (5)$\Rightarrow$(4). Before the proof, let us first notice that $f^{\ast}(y)=\bar{f}^{\ast}(y)$ for all $y\in L^{1}_{\mathcal{F}}(\mathcal{E})$: since $f^{\ast}(y)=\bigvee\{E(xy|\mathcal{F})-f(x)~|~x\in L^{\infty}(\mathcal{E})\}$, $\bar {f}^{\ast}(y)=\bigvee\{E(xy|\mathcal{F})-\bar{f}(x)~|~x\in L^{\infty}_{\mathcal{F}}(\mathcal{E})\}$ and $L^{\infty}_{\mathcal{F}}(\mathcal{E})=H_{cc}(L^{\infty}(\mathcal{E}))$ by Lemma 4.22, then $\bar{f}^{\ast}(y)=\bigvee\{E(xy|\mathcal{F})-f(x)~|~x\in L^{\infty}(\mathcal{E})\}=f^{\ast}(y)$ by noticing that both $E(xy|\mathcal{F})$ and $\bar{f}(x)$ are local functions of $x\in L^{\infty}_{\mathcal{F}}(\mathcal{E})$ for each fixed $y\in L^{1}_{\mathcal{F}}(\mathcal{E})$ and applying (2) of Lemma 3.16 to the local function $g:L^{\infty}_{\mathcal{F}}(\mathcal{E})\rightarrow \bar{L}^{0}(\mathcal{F})$ defined by $g(x)=E(xy|\mathcal{F})-\bar{f}(x)$. (4)$\Rightarrow$(5). by (4) of Lemma 3.16 it is clear that $\bar{g}:L^{\infty}_{\mathcal{F}}(\mathcal{E})\rightarrow \bar{L}^{0}(\mathcal{F})$ defined by $\bar{g}(x)=\bigvee\{E(xy|\mathcal{F})-\bar{f}^{\ast}(y)~|~y\in L^{1}_{\mathcal{F}}(\mathcal{E}),y\leq 0$ and $E(y|\mathcal{F})=-1\}$ for any $x\in L^{\infty}_{\mathcal{F}}(\mathcal{E})$ is local since $E(xy|\mathcal{F})-\bar{f}^{\ast}(y)$ is a local function of $x$ for each fixed $y$. By applying (3) of Lemma 3.16 to $\bar{f}$ and $\bar{g}$ one can see that $\bar{f}=\bar{g}$ since $L^{\infty}_{\mathcal{F}}(\mathcal{E})= H_{cc}(L^{\infty}(\mathcal{E}))$ and it is just (4) that $\bar{f}(x)=\bar{g}(x)$ for all $x\in L^{\infty}(\mathcal{E})$. (5)$\Rightarrow$(6) is clear. (6)$\Rightarrow$(5) is by Theorem 4.17. \hfill $\square$ \end{proof} \subsection*{4.4 Extension theorem for $L^p$--conditional risk measures} Motivated by the work in \cite{Zowe,SGP}, D. Filipovi\'{c}, M. Kupper and N. Vogelpoth introduced the following $L^p$--conditional risk measures in \cite{FKV-appro} as follows: \begin{definition} (See \cite{FKV-appro}.) Let $1\leq r\leq p<+\infty$ and $f$ be a function from $L^p(\mathcal{E})$ to $L^{r}(\mathcal{F})$. $f$ is an $L^p$--conditional risk measure if the following two conditions are satisfied: \noindent (1). $f(x)\leq f(y)$ for all $x,y\in L^p(\mathcal{E})$ with $x\geq y$; \noindent (2). $f(x+y)=f(x)-y$ for all $x\in L^p(\mathcal{E})$ and $y\in L^p(\mathcal{F})$. \end{definition} The following representation result is known: \begin{proposition} ($See$ \cite{FKV-appro}.) Let $f$ be an $L^{0}(\mathcal{F})$--convex $L^p$--conditional risk measure from $L^p(\mathcal{E})$ to $L^{r}(\mathcal{F})$ and $1\leq r\leq p<+\infty$. If $f$ is continuous from $(L^p(\mathcal{E}),\|\cdot\|_p)$ to $(L^r(\mathcal{E}),\|\cdot\|_r)$, then $f(x)=\bigvee\{E(xy|\mathcal{}F)-f^{\ast}(y)~|~y\in L^{q}(\mathcal{E}),y\leq 0,E(|y|^{q}|\mathcal{F})\in L^{\frac{r(p-1)}{p-r}}(\mathcal{F})$ and $E(y|\mathcal{F})=-1\}$ for all $x\in L^p(\mathcal{E})$, where $\frac{1}{p}+\frac{1}{q}=1$ and $\frac{r(p-1)}{p-r}=+\infty$ when $p=r$ and $f^{\ast}:L^{q}_{\mathcal{F}}(\mathcal{E})\rightarrow \bar{L}^0(\mathcal{F})$ is defined by $f^{\ast}(y)=\bigvee\{E(xy|\mathcal{F})-f(x)~|~x\in L^p(\mathcal{E})\}$ for all $y\in L^{q}_{\mathcal{F}}(\mathcal{E})$. \end{proposition} The aim of this subsection is to give an extension theorem for any $L^0(\mathcal{F})$--convex $L^p$--conditional risk measure, in particular, in this process we also improve Proposition 4.27 in that we can give a sufficient and necessary condition that any $L^0(\mathcal{F})$--convex $L^p$--conditional risk measure can be represented as in Proposition 4.27, in fact, a new and shorter proof of Proposition 4.27 will be also given. An important special case of $L^0(\mathcal{F})$--convex $L^p$--conditional risk measures is the following conditional risk measure derived from the solution of backward stochastic differential equations (BSDE, for short). Let $(B_t)_{t\geq 0}$ be a standard $d$--dimensional Brown motion defined on a probability space $(\Omega,\mathcal{F},P)$ and $(\mathcal{F}_t)_{t\geq 0}$ the augmented filtration generated by $(B_t)_{t\geq 0}$. Let a function $g:R^{+}\times\Omega\times R^d\rightarrow R$, $(t,\omega,z)\rightarrow g(t,\omega,z)$ (briefly,$g(t,z)$) satisfy the following conditions: \noindent (A). $g$ is Lipschitz in $z$, i.e., there exists a constant $\mu>0$ such that we have $dt\times dP$--a.s., for any $z_0,z_1\in R^d$, $|g(t,z_0)-g(t,z_1)|\leq \mu\|z_0-z_1\|$. \noindent (B). For all $z\in R^d$, $g(\cdot,z)$ is a predictable process such that for any $T>0$, $E[\int^{T}_{0}g(t,\omega,z)^2 dt]<+\infty$ for any $z\in R^d$. \noindent (C). $dt\times dP$--a.s., $g(t,0)=0$. \noindent (D). $g$ is convex in $z$: $\forall \alpha\in[0,1]$, $\forall z_0,z_1\in R^d$, $dt\times dP$--a.s., $g(t,\alpha z_0+(1-\alpha)z_1)\leq \alpha g(t,z_0)+(1-\alpha)g(t,z_1)$. Then, for any $T>0$, the following BSDE: \[ \left\{ \begin{array}{ll} -dY_t=g(t,Z_t)dt-Z_tdB_t, \\ Y_T=\xi, \end{array} \right. \]\\ \noindent where $\xi\in L^{2}(\Omega,\mathcal{F}_T,P)$, has a unique solution $(Y_t,Z_t)_{t\in[0,T]}$ consisting of predictable stochastic processes such that $E[\int^{T}_{0}Y_t^{2}dt]<+\infty$ and $E[\int^{T}_{0}\|Z_t\|^{2}dt]<+\infty$, cf. \cite{DPG,TXG-dual}. Peng defined the conditional $g$--expectation of $\xi$ at time $t$ as $$\mathcal{E}_g(\xi|\mathcal{F}_t):=Y_t.$$ Now, for any fixed $t\in[0,T]$, define $\rho^{g}_{t}(\cdot):L^2(\mathcal{F}_T)\rightarrow L^2(\mathcal{F}_t)$ by $\rho^{g}_{t}(x)=\mathcal{E}_g(-x|\mathcal{F}_t)$ for all $x\in L^2(\mathcal{F}_T)$, then $\rho^{g}_{t}$ is an $L^{0}(\mathcal{F}_t)$--convex $L^2$--conditional risk measure. By Theorem 3.2 of \cite{SGP}, $\rho^{g}_{t}$ is continuous from $(L^2(\mathcal{F}_T),\|\cdot\|_2)$ to $(L^2(\mathcal{F}_t),\|\cdot\|_2)$. Again by Theorem 3.2 of \cite{SGP}, $|\rho^{g}_{t}(x)-\rho^{g}_{t}(y)|\leq c(E[|x-y|^2|\mathcal{F}_t])^{\frac{1}{2}}$ for all $x,y\in L^2(\mathcal{F}_T)$, where $c=e^{8(1+\mu^2)(T-t)}$, namely when $L^2(\mathcal{F}_T)$ is regarded as a subspace of the $RN$ module $(L^2_{\mathcal{F}_t}(\mathcal{F}_T),|||\cdot|||_2)$ and $L^2(\mathcal{F}_t)$ is regarded as a subspace of the $RN$ module $(L^0(\mathcal{F}_t),|\cdot|)$, $\rho^{g}_t$ is Lipschitz with respect to the $L^0$--norms. So, $\rho^{g}_t$ can be uniquely extended to an $L^0(\mathcal{F}_t)$--convex $L^2_{\mathcal{F}_t}(\mathcal{F}_T)$--conditional risk measure $\bar{\rho}^{g}_t$ such that $|\bar{\rho}^{g}_{t}(x)-\bar{\rho}^{g}_{t}(y)|\leq c(E[|x-y|^2|\mathcal{F}_t])^{\frac{1}{2}}$ for all $x,y\in L^2_{\mathcal{F}_t}(\mathcal{F}_T)$ since $L^2(\mathcal{F}_T)$ is $\mathcal{T}_{\varepsilon,\lambda}$--dense in $(L^2_{\mathcal{F}_t}(\mathcal{F}_T),|||\cdot|||_2)$ (note: $L^2(\mathcal{F}_T)=L^2(L^2_{\mathcal{F}_t}(\mathcal{F}_T))$ and use Theorem 4.20), the proof is the same as that of Theorem 4.21. Since a general $L^0(\mathcal{F})$--convex $L^p$--conditional risk measure from $L^p(\mathcal{E})$ to $L^r(\mathcal{F})$ may not necessarily $\mathcal{T}_{\varepsilon,\lambda}$--uniformly continuous when $L^p(\mathcal{E})$ is regarded a subspace of $(L^p_{\mathcal{F}}(\mathcal{E}),|||\cdot|||_p)$ and $L^{r}(\mathcal{F})$ as a subspace of $(L^0(\mathcal{F}),|\cdot|)$, we are forced to use a constructive way to obtain a unique extension, namely Theorem 4.28 below. \begin{theorem} Let $f:L^p(\mathcal{E})\rightarrow L^r(\mathcal{F})$ be an $L^0(\mathcal{F})$--convex $L^p$--conditional risk measure. Then there is a unique $L^0(\mathcal{F})$--convex $L^p_{\mathcal{F}}(\mathcal{E})$--conditional risk measure $\bar{f}:L^p_{\mathcal{F}}(\mathcal{E})\rightarrow L^0(\mathcal{F})$ such that $\bar{f}|_{L^{p}(\mathcal{E})}=f$. \end{theorem} \begin{proof} Define $\bar{f}:L^p_{\mathcal{F}}(\mathcal{E})\rightarrow L^0(\mathcal{F})$ by $\bar{f}(x)=\Sigma_{n=1}^{\infty}\tilde{I}_{A_n}f(x_n)$ for any canonical representation $\Sigma_{n=1}^{\infty}\tilde{I}_{A_n}x_n$ of $x$, First, $\bar{f}$ is well defined. In fact, for any two canonical representations $\Sigma_{n=1}^{\infty}\tilde{I}_{A_n}x_n$ and $\Sigma_{n=1}^{\infty}\tilde{I}_{B_n}y_n$ of $x$, $\Sigma_{n=1}^{\infty}\tilde{I}_{A_n}f(x_n)=\Sigma_{i,j=1}^{\infty}\tilde{I}_{A_i\bigcap B_j}f(x_i)=\Sigma_{i,j=1}^{\infty}$ $\tilde{I}_{A_i\bigcap B_j}f(y_j)=\Sigma_{n=1}^{\infty}\tilde{I}_{B_n}f(y_n)$. Second, $\bar{f}$ is $L^0(\mathcal{F})$--convex: let $\xi\in L^0_+(\mathcal{F})$ such that $0\leq\xi\leq1$ and $x,y\in L^p_{\mathcal{F}}(\mathcal{E})$ have the canonical representations $\Sigma_{n=1}^{\infty}\tilde{I}_{A_n}x_n$ and $\Sigma_{n=1}^{\infty}\tilde{I}_{B_n}y_n$, respectively. Then $x=\Sigma_{i,j=1}^{\infty}\tilde{I}_{A_i\bigcap B_j}x_i$ and $y=\Sigma_{i,j=1}^{\infty}\tilde{I}_{A_i\bigcap B_j}y_j$, and thus $\bar{f}(\xi x+(1-\xi)y)=\bar{f}(\Sigma_{i,j=1}^{\infty}\tilde{I}_{A_i\bigcap B_j}(\xi x_i+(1-\xi)y_j))=\Sigma_{i,j=1}^{\infty}\tilde{I}_{A_i\bigcap B_j}f(\xi x_i+(1-\xi)y_j)\leq \xi\Sigma_{i,j=1}^{\infty}\tilde{I}_{A_i\bigcap B_j}f(x_i)+(1-\xi)\Sigma_{i,j=1}^{\infty}\tilde{I}_{A_i\bigcap B_j}f(y_j)=\xi \bar{f}(x)+(1-\xi)\bar{f}(y)$. Similarly, $\bar{f}$ is also monotone. Further, by Lemma 4.22 and the local property of $\bar{f}$, $\bar{f}$ is also cash invariant in the sense of Definition 4.15. Finally, any $L^0(\mathcal{F})$--convex $L^p_{\mathcal{F}}(\mathcal{E})$--conditional risk measure $g$ with $g|_{L^p(\mathcal{E})}$ $=f$ must coincide with $\bar{f}$ since $g$ has the local property by $L^{p}_{\mathcal{F}}(\mathcal{E})=H_{cc}(L^p(\mathcal{E}))$ and applying (3) of Lemma 3.16 to $\bar{f}$ and $g$, which proves the uniqueness. \hfill $\square$ \end{proof} Theorem 4.31 below shows that the continuity of $f$ in Proposition 4.27 can be weakened to that $\bar{f}$ is $\mathcal{T}_{\varepsilon,\lambda}$ (or equivalently, $\mathcal{T}_{c}$)-lower semicontinuous, whereas the implication of the continuity of $f$ is reflected to some extent by Theorem 4.29 below. \begin{theorem} Let $f$ and $\bar{f}$ be the same as in Theorem 4.28. If $f$ is continuous from $(L^p(\mathcal{E}),\|\cdot\|_p)$ to $(L^r(\mathcal{F}),\|\cdot\|_r)$, then $\bar{f}$ is $\mathcal{T}_{\varepsilon,\lambda}$--continuous from $(L^p_{\mathcal{F}}(\mathcal{E}),|||\cdot|||_p)$ to $(L^0(\mathcal{F}),|\cdot|)$. \end{theorem} \begin{proof} When $L^p(\mathcal{E})$ is regard as a subspace $(L^p_{\mathcal{F}}(\mathcal{E}),|||\cdot|||_p)$ and $(L^r(\mathcal{F}),\|\cdot\|_r)$ is regarded as a subspace of $(L^0(\mathcal{F}),|\cdot|)$, we first prove that $f$ is $\mathcal{T}_{\varepsilon,\lambda}$--continuous from $(L^p(\mathcal{E}),|||\cdot|||_p)$ to $(L^r(\mathcal{F}),|\cdot|)$. For this, we only need to prove that , for each fixed $x_{0}\in L^p(\mathcal{E})$ and each sequence $\{x_n,n\in N\}$ in $L^p(\mathcal{E})$ such that $\{E[|x_n-x_{0}|^p|\mathcal{F}]^{\frac{1}{p}}:n\in N\}$ converges in probability measure $P$ to 0, there exists a subsequence $\{x_{n_k},k\in N\}$ of $\{x_n,n\in N\}$ such that $\{f(x_{n_k}),k\in N\}$ converges in probability measure $P$ to $f(x_{0})$. Since $f$ is $\mathcal{F}$--local, we only need to prove that, for any positive number $\delta$, there exist an $\mathcal{F}$--measurable subset $H_{\delta}$ of $\Omega$ and a subsequence $\{x_{n_k},k\in N\}$ of $\{x_n,n\in N\}$ such that $P(\Omega\setminus H_{\delta})>1-\delta$ and $\{f(x_{n_k}),k\in N\}$ converges in probability measure $P$ to $f(x_{0})$ on $\Omega\setminus H_{\delta}$. In fact, by the Egoroff theorem there are such $H_{\delta}$ and $\{x_{n_k},k\in N\}$ such that $\{E[|x_{n_k}-x_{0}|^{p}|\mathcal{F}]^{\frac{1}{p}},k\in N\}$ converges uniformly to 0 on $\Omega\setminus H_{\delta}$, so that $\{\tilde{I}_{\Omega\setminus H_{\delta}}x_{n_k},k\in N\}$ converges to $\tilde{I}_{\Omega\setminus H_{\delta}}x_{0}$ in the usual $L^p$--norm $\|\cdot\|_p$ by the Lebesgue domination convergence theorem, hence $\{\tilde{I}_{\Omega\setminus H_{\delta}}f(x_{n_k}),k\in N\}$ converges in the $L^r$--norm to $\tilde{I}_{\Omega\setminus H_{\delta}}f(x_{0})$, which implies that $\{f(x_{n_k}),k\in N\}$ converges in probability measure $P$ to $f(x_{0})$ on $\Omega\setminus H_{\delta}$. We can now prove that $\bar{f}$ is $\mathcal{T}_{\varepsilon,\lambda}$--continuous. Let $\{x_k,k\in N\}$ be a sequence in $L^p_{\mathcal{F}}(\mathcal{E})$ convergent in $\mathcal{T}_{\varepsilon,\lambda}$ to $x\in L^p_{\mathcal{F}}(\mathcal{E})$, where $\mathcal{T}_{\varepsilon,\lambda}$ is the $(\varepsilon,\lambda)$--topology on $L^p_{\mathcal{F}}(\mathcal{E})$ induced by $|||\cdot|||_p$, then for any canonical representation $\Sigma_{n=1}^{\infty}\tilde{I}_{A_n}x_n$ of $x$, we only need to prove that $\{\bar{f}(x_k),k\in N\}$ converges in probability $P$ to $\bar{f}(x)$ on each fixed $A_n$. Now, let $n_0\in N$ be fixed. For each given canonical representation $\Sigma_{n=1}^{\infty}\tilde{I}_{A_n^k}x_n^k$ of $x_k$, we choose $m_k\in N$ such that $P(\Sigma_{n\geq m_k} A_n^k)<\frac{1}{k}$, then $\{\Sigma_{n=1}^{m_k}\tilde{I}_{A_n^k}x_n^k,k\in N\}$ still converges in $\mathcal{T}_{\varepsilon,\lambda}$ to $x$, hence $\{\tilde{I}_{A_{n_0}}\Sigma_{n=1}^{m_k}\tilde{I}_{A_n^k}x_n^k,k\in N\}$, of course, converges in $\mathcal{T}_{\varepsilon,\lambda}$ to $\tilde{I}_{A_{n_0}}x (=\tilde{I}_{A_{n_0}}x_{n_0})$. By what we have proved, $\{f(\tilde{I}_{A_{n_0}}\Sigma_{n=1}^{m_k}\tilde{I}_{A_n^k}x_n^k),k\in N\}$ converges in the probability measure $P$ to $f(\tilde{I}_{A_{n_0}}x_{n_0})$. By the $\mathcal{F}$--local property of $f$, $\tilde{I}_{A_{n_0}}f(\Sigma_{n=1}^{m_k}\tilde{I}_{A_n^k}x_n^k)=\tilde{I}_{A_{n_0}}f(\tilde{I}_{A_{n_0}}\Sigma_{n=1}^{m_k}\tilde{I}_{A_n^k}x_n^k)$ and $\tilde{I}_{A_{n_0}}f(x_{n_0})=\tilde{I}_{A_{n_0}}f(\tilde{I}_{A_{n_0}}x_{n_0})$, then $\{f(\tilde{I}_{A_{n_0}}\Sigma_{n=1}^{m_k}\tilde{I}_{A_n^k}x_n^k),k\in N\}$ converges in the probability measure $P$ to $f(x_{n_0})$ on $A_{n_0}$. Finally, since $\bar{f}(x_k)=\tilde{I}_{\bigcup_{n=1}^{m_k}A_{n}^{k}}f(\Sigma_{n=1}^{m_k}\tilde{I}_{A_n^k}x_n^k))+\Sigma_{n\geq m_k}^{\infty}\tilde{I}_{A_n^k}f(x_n^k)$ and $\tilde{I}_{A_{n_0}}$ $f(x)=\tilde{I}_{A_{n_0}}\bar{f}(x)=\tilde{I}_{A_{n_0}}\bar{f}(\tilde{I}_{A_{n_0}}x)=\tilde{I}_{A_{n_0}}f(x_{n_0})$, we have that $\{\bar{f}(x_k),k\in N\}$ converges in the probability measure $P$ to $\bar{f}(x)$ on $A_{n_0}$ by noticing that $P(\Sigma_{n\geq m_k}^{\infty}A_n^k)\rightarrow 0$ when $k\rightarrow \infty$. \hfill $\square$ \end{proof} \begin{lemma} Let $f:L^p({\mathcal{E}})\rightarrow L^r({\mathcal{F}})$ be an $L^0(\mathcal{F})$--convex $L^p$--conditional risk measure and $\bar{f}:L^{p}_{\mathcal{F}}(\mathcal{E})\rightarrow L^0(\mathcal{F})$ its unique extension. Then we have that $f^{\ast}(y)=\bar{f}^{\ast}(y)$ for all $y\in L^{q}_{\mathcal{F}}(\mathcal{E})$. \end{lemma} \begin{proof} let us recall: $f^{\ast}(y)=\bigvee\{E(xy|\mathcal{F})-f(x)~|~x\in L^p(\mathcal{E})\}$ and $\bar{f}^{\ast}(y)=\bigvee\{E(xy|\mathcal{F})-\bar{f}(x)~|~x\in L^p_{\mathcal{F}}(\mathcal{E})\}$. By Lemma 4.22, $L^p_{\mathcal{F}}(\mathcal{E})=H_{cc}(L^p(\mathcal{E}))$, which, together with the local property of $g:L^{p}_{\mathcal{F}}(\mathcal{E})\rightarrow \bar{L}^{0}(\mathcal{F})$ defined by $g(x)=E(xy|\mathcal{F})-\bar{f}(x)$ for all $x\in L^{p}_{\mathcal{F}}(\mathcal{E})$ implies the $f^{\ast}(y)=\bar{f}^{\ast}(y)$ for all $y\in L^q_{\mathcal{F}}(\mathcal{E})$ by applying (2) of Lemma 3.16 to the local function $g$ and $G:=L^{p}(\mathcal{E})$. \hfill $\square$ \end{proof} \begin{theorem} Let $1\leq r\leq p<+\infty$, $q$ be the H\"{o}lder conjugate number of $p$, $f:L^p({\mathcal{E}})\rightarrow L^r({\mathcal{F}})$ an $L^0(\mathcal{F})$--convex $L^p$--conditional risk measure and $\bar{f}:L^{p}_{\mathcal{F}}(\mathcal{E})\rightarrow L^0(\mathcal{F})$ the unique extension of $f$ determined by Theorem 4.28. Then the following statements are equivalent: \noindent $(1)$. $f(x)=\bigvee\{E(xy|\mathcal{F})-f^{\ast}(y)~|~y\in L^{q}(\mathcal{E}),y\leq 0,E(|y|^{q}|\mathcal{F})\in L^{\infty}(\mathcal{F})$ and $E(y|\mathcal{F})=-1\}$ for all $x\in L^{p}(\mathcal{E})$. \noindent $(2)$. For any given positive number $\gamma$, $f(x)=\bigvee\{E(xy|\mathcal{F})-f^{\ast}(y)~|~y\in L^{q}(\mathcal{E}),y\leq 0,E(|y|^{q}|\mathcal{F})\in L^{\gamma}(\mathcal{F})$ and $E(y|\mathcal{F})=-1\}$ for all $x\in L^{p}(\mathcal{E})$. \noindent $(3)$. $f(x)=\bigvee\{E(xy|\mathcal{F})-f^{\ast}(y)~|~y\in L^{q}(\mathcal{E}),y\leq 0,E(|y|^{q}|\mathcal{F})\in L^{\frac{r(p-1)}{p-r}}(\mathcal{F})$ and $E(y|\mathcal{F})=-1\}$ for all $x\in L^{p}(\mathcal{E})$. \noindent $(4)$. $f(x)=\bigvee\{E(xy|\mathcal{F})-f^{\ast}(y)~|~y\in L^{q}_{\mathcal{F}}(\mathcal{E}),y\leq 0$ and $E(y|\mathcal{F})=-1\}$ for all $x\in L^{p}(\mathcal{E})$. \noindent $(5)$. $\bar{f}(x)=\bigvee\{E(xy|\mathcal{F})-\bar{f}^{\ast}(y)~|~y\in L^{q}_{\mathcal{F}}(\mathcal{E}),y\leq 0$ and $E(y|\mathcal{F})=-1\}$ for all $x\in L^{p}_{\mathcal{F}}(\mathcal{E})$. \noindent $(6)$. $\bar{f}$ is $\mathcal{T}_{\varepsilon,\lambda}$ (or equivalently, $\mathcal{T}_{c}$)-lower semicontinuous. \noindent $(7)$. $\bar{f}$ is continuous from $(L^{p}_{\mathcal{F}}(\mathcal{E}),\mathcal{T}_c)$ to $ (L^{0}(\mathcal{F}),\mathcal{T}_c)$. \end{theorem} \begin{proof} For any fixed $x\in L^{p}(\mathcal{E})$, let $g(y)=E(xy|\mathcal{F})-f^{\ast}(y)$ for any $y\in L^{q}_{\mathcal{F}}(\mathcal{E})$, then $g$ has the local property. By applying (2) of Lemma 3.16 and (2) of Lemma 4.24 one can easily see that (1)$\Leftrightarrow$(2)$\Leftrightarrow$(3)$\Leftrightarrow$(4). By Lemma 4.30, $f^{\ast}=\bar{f}^{\ast}$, so (5)$\Rightarrow$(4) is clear. If (4) is true, let $\bar{g}:L^{p}_{\mathcal{F}}(\mathcal{E})\rightarrow \bar{L}^{0}(\mathcal{F})$ be defined by $\bar{g}(x)=\bigvee\{E(xy|\mathcal{F})-\bar{f}^{\ast}(y)~|~y\in L^{q}_{\mathcal{F}}(\mathcal{E}),y\leq 0$ and $E(y|\mathcal{F})=-1\}$ for any $x\in L^{p}_{\mathcal{F}}(\mathcal{E})$, then by (3) of Lemma 3.16 we have that $\bar{f}=\bar{g}$ since (4) just shows that $\bar{f}(x)=\bar{g}(x)$ for any $x\in L^p(\mathcal{E})$, which shows (4)$\Rightarrow$(5). (5)$\Rightarrow$(6) is clear. (6)$\Rightarrow$(5) is by Theorem 4.16. (6)$\Rightarrow$(7) is by Theorem 3.44. (7)$\Rightarrow$(6) is clear. \hfill $\square$ \end{proof} Theorem 4.31 not only gives a very short proof of D. Filipovi\'{c}, M. Kupper and N. Vogelpoth's Proposition 4.27, whose original proof in \cite{FKV-appro} is somewhat complicated, but also improves Proposition 4.27 in that we give a sufficient and necessary condition for (3) of Theorem 4.31 to hold, namely that $\bar{f}$ is $\mathcal{T}_{\varepsilon,\lambda}$ (or $\mathcal{T}_{c}$)-lower semicontinuous. Besides, (1) of Theorem 4.31 maybe is more interesting. The whole Section 3 together with Theorems 4.5 and 4.13 and Corollary 4.14 has formed a complete random convex analysis and thus this paper has provided a solid analytic foundation for the module approach to conditional risk measures. Furthermore, Theorems 4.25 and 4.31 not only give the complete relations among three kinds of conditional convex risk measures, which shows that $L^p$--conditional risk measures can be incorporated into $L^{p}_{\mathcal{F}}(\mathcal{E})$--conditional risk measures ($1\leq p\leq+\infty$), but also their proofs provide many useful analytic skills. Maybe the module approach together with these skills will develop their power in the future study of dynamic risk measures with the model spaces consisting of stochastic processes, this topic just began in \cite{CDK}. Just as classical convex analysis has many other rich applications as well as the application to convex risk measures, cf. \cite{ET}, we may also hope that random convex analysis can be applied in other aspects. \section*{Acknowledgements} The first author of this paper thanks Professor Zhenting Hou for invaluable talks and encouragement and Professor Quanhua Xu for kindly providing us the excellent literature \cite{HLR} in February, 2012 to make us know, for the first time, the existence of \cite{HLR} since \cite{HLR} has never been mentioned before in the literature of related fields (e.g. \cite{SS}).
\section{Introduction} \label{sec:intro} Starting from the solution of the non linear heat equation, Peng \cite{P1,P2} has introduced the notion of $G$-expectation $({\cal E}^G_t)_{t \in \R_+}$ which is a sublinear time consistent dynamic procedure defined on continuous functions on $\Omega$. The state space $\Omega$ is there equal to the set of continuous paths, and ${\cal E}^G_t(X)$ is defined by a stepwise evaluation of the partial differential equation (PDE). Denis at al \cite{DHP} have then proved that ${\cal E}^G_0(X)= \sup_{P^{\theta} \in {\cal P}}E_{P^{\theta}}(X)$, where ${\cal P}$ is a weakly compact set of probability measures which are non dominated (i.e. there is no probability measure $P$ such that $Q \ll P$ for all $Q$ in ${\cal P}$). The properties of $({\cal E}^G_t)_{t \in \R_+}$ are up to a minus sign the properties of a sublinear time consistent dynamic risk measure. Since the seminal papers of Peng on $G$-expectation, a new challenge is to develop a theory of time consistent dynamic procedures in a non dominated framework, and also to provide viscosity solutions to fully non linear second order partial differential equations.\\ Some recent works study the properties of these procedures, in the static case\cite{BNK1} and in the dynamic case \cite{NS,BNK2}. Other works construct time consistent dynamic procedures. The first construction of $G$-expectation by Peng \cite{P1,P2} was made on the state of continuous paths from the non linear heat equation $-\partial_t v(t,x)-G(D^2v(t,x))=0$, $G(A)=\sup_{\gamma\in {\Theta}}Tr(\gamma \gamma^*A)$ ($\Theta$ bounded) with boundary condition $v(0,x)=f(x)$. Starting from the unique viscosity solution of a second order PDE containing a non local term, this construction has been extended in \cite{HP} to produce a time consistent sublinear procedure on the set of c\`adl\`ag paths. In both cases $G$ is independent of $t$ and $x$ and the $G$-heat equation corresponds to diffusions with no drift and a diffusion coefficient varying between two bounds. These works of Peng {\it et al.} have motivated other works constructing time consistent dynamic procedures on the set of continuous paths, in a non dominated framework. Nutz has produced time consistent sublinear procedures allowing the $G$ coefficient to vary with both $t$ and $x$ (or equivalently with $\omega$ in the non Markovian case) in \cite{N1}, or allowing both the diffusion coefficient and the drift term to vary between bounds \cite{N2}. In both cases the construction is made $\omega$ by $\omega$. Starting from a given set of probability measures for all $t$ and $\omega$, satisfying some compatibility conditions and analytic properties, Nutz and van Handel \cite{NVH} construct also sublinear time consistent procedures on the space of continuous paths. Soner et al \cite{STZ1,STZ2} have constructed convex (and not only sublinear) time consistent dynamic procedures in the setting of diffusions as well. Their approach makes use of a solution of a Backward Stochastic Differential Equation (BSDE) associated to every probability measure in the set ${\cal P}$ and then solves an ``aggregation problem''. The diffusion coefficient varies in a domain independent of $t$ and $x$. In \cite{STZ2} it is proved that this construction gives rise to viscosity solutions to fully non linear second order PDE.\\ Time consistent dynamic procedures on a filtered probability space or up to a minus sign dynamic risk measures have been studied in many papers \cite{BN03,BN04,CDK,D,KS}. Risk measures are characterized by their dual representation. In particular, in the static case \cite{FS,FR}, $\rho_{0,t}(X) =\sup_{Q \in {\cal P}}(E_Q(-X)-\alpha(Q))$ where ${\cal P}$ is a set of probability measures all absolutely with respect to $P$. Sublinear time consistent risk measures are fully characterized by a stable set of probability measures \cite{D}. Convex time consistent dynamic risk measures are described by a stable set of probability measures ${\cal P}$ and a local penalty defined on ${\cal P}$ satisfying the cocycle condition \cite{BN03,BN04}. In the particular case of the Brownian filtration, time consistent dynamic risk measures are limits of solutions of BSDE \cite{DPR}.\\ In order to generalize the construction of time consistent dynamic procedures to the case of non dominated probability measures, the main point is to understand the notion of stable set of probability measures in this new framework. Indeed the usual notion of stable set involves $Q$-conditional expectation. However $Q$-conditional expectation is defined up to a $Q$-null set and this is a drawback when one considers non equivalent probability measures. On a Polish space $\Omega$, for all sub $\sigma$- algebra ${\cal B}$ of the Borel $\sigma$ algebra, there exists always a measurable version of the $Q$-condition expectation called regular conditional probability distribution given ${\cal B}$ (\cite{SV4}, Theorem 1.1.6.). However there is no unicity of the $Q$-regular conditional distribution given ${\cal B}$. In order to extend the notion of stability by bifurcation \cite{BN03} to a set of non dominated probability measures as well as the local condition for a penalty \cite{BN03}, one needs to be able to make a coherent choice for a measurable version of the conditional expectation.\\ The first goal of the present paper is to prove that weakly continuous solutions to a martingale problem (Section\ref{secMARPB}) is a very nice setting in which there exists a canonical version of the conditional expectation (Section \ref{seccano}). This allows to extend the notion of stability to a set of probability measures which are non dominated (Section \ref {secstable}). This canonical version has furthermore continuity properties (Section \ref{secconti}). Specializing to the case of continuous diffusions or to the case of diffusions with Levy generator, we then construct penalties having a Feller property (Section \ref{secTCCC}), satisfying the local condition and the cocycle condition. The canonical version of conditional expectation joint with the Feller penalties give a generic and constructive method to produce ${\cal T}$-time consistent convex dynamic procedures on the space of continuous paths and also on the space of c\`adl\`ag paths in a non dominated framework. The set ${\cal T}$ can be any subset of $\R^*_+$ in particular it can be $\R^*_+$ or a discrete subset of $\R^*_+$. This construction generalizes to the non dominated framework the construction of \cite{JBN-PDE}. It can be used to construct a great variety of time consistent dynamic procedures in non dominated framework. In the case of continuous paths, the underlying set of probability measures ${\cal P}$ is a ``stable set'' generated by probability measures solution to a martingale problem associated to a continuous diffusion with diffusion coefficient $a(t,x)$ and drift coefficient $b(t,x)$ continuous bounded , $a(t,x)$ invertible for all $(t,x)$ such that $(a,b)$ takes values in a multivalued Borel set. In the case of c\`adl\`ag paths, the underlying set of probability measures ${\cal P}$ is the ``stable set'' generated by probability measures solutions to a martingale problem associated to diffusions with Levy generator with coefficients $a(t,x)$ $b(t,x)$ satisfying the above conditions and the jump measure $M(t,x)$ satisfying the hypothesis of \cite{St}, $(a,b,M)$ taking values in a multivalued Borel set. The procedure is defined on the closure (for the norm $\sup_{P \in {\cal P}}E_P(|X|)$ of the lattice vector space of continuous coordinate functions $f(X_{t_1},..X_{t_k})$ ($f$ continuous on $(\R^n)^k$) and also on the cone $f(X_{t_1},..X_{t_k})$, where $f$ is lower semi continuous bounded from below. A construction for a time consistent procedure on c\`adl\`ag paths is independently proposed in \cite{KPZ}, but within a very limited and specific framework: the set of probability mesasures is generated by probability measures solutions to the martingale problem associated to diffusions with Levy generator where there is no drift, the diffusion coefficient and the jump measure depend only of $t$ (not on $x$), there is no multivalued Borel mapping, and no penalty function. Furthermore the procedure is defined on the uniformly continuous functions on the space of c\`adl\`ag paths, which does not allow to study PDEs in the context of c\`adl\`ag paths. \\ Our last goal is to prove as in \cite{JBN-PDE} that these procedures give a new probabilistic approach to second order PDE (Section \ref{viscsol}). Indeed, in the case ${\cal T}=\R_+$, these procedures applied to the random variable $h(X_t)$ ($h$ lower semi-continuous bounded from below) give rise to a time consistent convex Feller process $\tilde h(s,X_s)$. The function $\tilde h$ is lower semi-continuous on $[0,t] \times \R^n$. Making use of the martingale property of the probability measures, we prove that this lower semi-continuous function is a viscosity supersolution of a second order PDE: \begin{equation} \left \{\parbox{12cm}{ \begin{eqnarray} -\partial_u v(u,x)-f(u, x,Dv(u,x),D^2v(u,x),\tilde Kv(u,x))&=&0 \nonumber \label{eqedp00}\\ v(t,x)& = & h(x)\nonumber \end{eqnarray} }\right. \end{equation} The non local term $\tilde Kv(u,x)(y)=v(u,x+y)-v(u,x)-\frac{y^*\nabla v(u,x)}{1+||y||^2}$ is specific of the case of diffusions with Levy generator. In case of continuous diffusions $f$ depends only on $(u, x,Dv(u,x),D^2v(u,x))$, the PDE is a fully non linear second order partial differential equation. In this last case, our results can be compared to those of \cite{STZ2}. In \cite{STZ2}, the function $f$ can also depend on $v$ but there are restrictive conditions on the dual of $f$ which are not needed in our approach. This is due to the fact that the construction in \cite{STZ2} relies on the existence of solutions to BSDE and also on the aggregation of these solutions. On the contrary in the present paper we have an explicit construction of the time consistent procedure. Furthermore the stability property of the set of probability measures and the conditions on the penalties imply the time consitency for the procedure (Section \ref{secTCC}). As mentioned in \cite{JBN-PDE}, the time consistency for the procedure corresponds to the usual dynamic programming principle. Notice also that our probabilistic approach can be considered as a general control problem, where the set of control is a set of non dominated probability measures. As in \cite{JBN-PDE}, under some conditions the function $\tilde h$ is continuous and is a viscosity solution of the PDE. \section{Martingale problem} \label{secMARPB} We fix a finite horizon $T$. In the following for $r \leq T$, the state space $\Omega^{r}$ is either the set ${\cal C}([r,T], \R^n)$ of continuous paths on $[r,T]$ endowed with the topology of uniform convergence, or the set ${\cal D}([r,T]$ of c\`adl\`ag paths with the Skorokod topology \cite{Bil}. Recall that $\Omega^r$ is a Polish space i.e. a complete separable metrizable space \cite{Bil}. \\In both cases $(X_t)_{t \in \R^+}$ denotes the coordinate process, and ${\cal B}^r_t$ is the $\sigma$-algebra generated by $\{X_u, \;r \leq u \leq t\} $. \subsection{Diffusions with continuous coefficients} \label{secdiff} As in \cite{JBN-PDE} we consider the martingale problem introduced in \cite{SV1}. In this subsection the state space is the set of continuous paths, $\Omega^r={\cal C}^r={\cal C}([r,T],\R^n)$. Let $a:[0,T] \times \R^n \rightarrow M_n(\R)$ and $b:[0,T] \times \R^n \rightarrow \R^n$ be continuous bounded maps such that for all $(t,x)$, $a(t,x)$ is definite positive (i.e. $a$ is strictly elliptic). For given $\theta \in \R^n$ let \begin{equation} Y^{a,b}_{r,t}(\theta)=exp \{\theta^*(X(t)-X(r))-\int_r^t \theta^*b(u,X(u)) du -\frac{1}{2} \int_s^t \theta^* a(u,X(u)) \theta du \} \label{eqmar} \end{equation} Following \cite{SV1} one says that the probability measure $Q^{a,b}_{r,y}$ on $(\Omega^r, {\cal B}^r)$ is a solution to the martingale problem (\ref{eqmar}) starting from $y$ at time $r$ if for all $\theta \in \R^n$, $(Y^{a,b}_{r,t}(\theta))_{t \geq r}$ is a martingale on $(\Omega^r, ({\cal B}^r)_t)_{t \geq r},Q^{a,b}_{r,y})$, and if $Q^{a,b}_{r,y}(\{X_r=y\})=1$. Denote ${\cal C}^{\infty}_c (\R^n)$ the set of ${\cal C}^{\infty}$ functions with compact support. Denote $L^{a,b}_t= \frac{1}{2}\sum_1^{n}a_{ij}(t,x) \frac{\partial^2}{\partial x_i \partial x_j} + \sum_1^{n}b_{i}(t,x)\frac{\partial}{\partial x_i}$. From \cite{SV3} Theorem 2.1, the above martingale problem is equivalent to \begin{equation} Z^{a,b}_{r,t}=f(X_t)-f(X_r)-\int_r^t L^{a,b}_u(f)(X_u)du \;\text{is a $Q^{a,b}_{r,y}$-martingale for all f in} \;{\cal C}^{\infty}_c (\R^n) \label{eqmart} \end{equation} The study of the martingale problem has been extended in \cite{St} to the case of diffusions with Levy generators. \subsection{Diffusions with Levy generators} \label{secDiLev} In this subsection $\Omega^{r}$ is equal to ${\cal D}^r={\cal D}([r,T],\R^n)$, the set of c\`adl\`ag paths endowed with the Skorokod topology \cite{Bil}. \begin{definition} Hypothesis (M)\\ A Borelian map $M$ defined on $\R_+ \times \R^n$ with values in the set of $\sigma$-finite measures on $\R^n-\{0\}$ satisfies hypothesis (M) if for all Borelian subset $\Delta$ of $\R^n-\{0\}$, \begin{equation} \int _{\Delta}\frac{y}{1+||y||^2}M(s,x,dy)\;\text{is continuous bounded} \label{eqbound} \end{equation} \label{defM} \end{definition} Let $K^M_t$ be the operator associated to $M$: \begin{equation} K^M_t(f)(x)=\int[f(x+y)-f(x)-\frac{y^*\nabla f(x)}{1+||y||^2}]M(t,x,dy) \label{eqM} \end{equation} Recall the following result from \cite{St}, Theorem 4.3. \begin{proposition} For all $a$ and $b$ continuous bounded and $a$ strictly elliptic, for all $M$ satisfying the hypothesis (M), for all $r$ in $\R_+$ and all $y$ in $\R^n$, there is a unique probability measure $Q^{a,b,M}_{r,y}$ on $({\cal D}^r,{\cal B}^r)$ solution to the martingale problem \begin{equation} Z^{a,b,M}_{r,t}=f(X_t)-f(X_r)-\int_r^t (L^{a,b}_u + K^M_u)(f)(X_u)du \;\;\forall f \in {\cal C}^{\infty}_c \R^n) \label{eqmart2} \end{equation} starting from $y$ at time $r$ i. e. such that $Q^{a,b,M}_{r,y}(\{X_r=y\})=1$ \label{propMP2} \end{proposition} \subsection{Weakly continuous solution to a martingale problem} \label{secWCM} Motivated by the above martingale problem both for continuous diffusions and for diffusions with Levy generators, we consider in the following solutions to a martingale problem in a general setting. A finite horizon $T$ is given. The state space can be either the set of continuous or of c\`adl\`ag paths. \begin{definition} For all $s \in [0,T[$ and $y \in \R^n$, let $Q_{s,y}$ be the unique solution to the martingale problem $Z$ on $(\Omega^s, {\cal B}^s_T)$, starting from $y$ at time $s$. ($Z$ means a whole family $(Z^i)_{i \in I}$. \begin{itemize} \item $(Q_{s,y})_{y \in \R^n}$ is weakly continuous\\ if $Q_{s,y}$ is a continuous function of $y$ for the weak topology. \item The martingale problem is additive if for all $i \in I$, for all $0 \leq r \leq s \leq t$, $Z^i_{r,s}$ is ${\cal B}^r_s$ measurable, $Z^i_{r,s}$ is a right continuous function of $s$, and \begin{equation} Z^i_{r,t}=Z^i_{r,s}+Z^i_{s,t}\; \label{eqadd} \end{equation} \item The martingale problem is bounded if for all $ 0 \leq r \leq s$, $Z^i_{r,t}$ is bounded. \end{itemize} \label{defFellprop} \end{definition} \begin{definition} Hypothesis $H_{\Theta}$ $ \Theta$ is a set of Borelian maps defined on $\R_+ \times \R^n$ with values in a topological space $E$ such that \begin{enumerate} \item for all $\theta \in \Theta$, for all $0 \leq r < t\leq T$ and for all $y$ in $\R^n$, there is a unique probability measure $Q^{\theta}_{r,y}$ on $(\Omega^r_t,{\cal B}^r_t)$ solution to the additive bounded martingale problem $Z^{\theta}$ starting from $y$ at time $r$. \item for all $\theta \in \Theta$, and $r <T$, $(Q^{\theta}_{r,y})_{y \in \R^n}$ is weakly continuous. \end{enumerate} \label{not01} \end{definition} Continuous diffusions and diffusions with Levy generators provide examples for $\Theta$: \begin{proposition} The notations are those of Section \ref{secdiff}. Let $a$ and $b$ be continuous bounded on $([0,T] \times \R^n)$ such that $a(s,x)$ is invertible for all $(s,x)$. Let $Z^{a,b}_{r,t} $ be given by equation (\ref{eqmart}). The unique solution $Q^{a,b}_{r,y}$ on $({\cal C}^r, {\cal B}^r_T)$ to the additive martingale problem $Z^{a,b}_{r,t}$ starting from $y$ at time $r$ is weakly continuous.\begin{equation} \Theta=\{(a,b) \;\text{continuous bounded,}\;a \;\text{invertible}\} \label{eqThetac} \end{equation} satisfies hypothesis $H_{\Theta}$. \label{propFell} \end{proposition} {\bf Proof} The weak continuity follows Theorem 7.1 of \cite{SV2}). See also Proposition 2.6 of \cite{JBN-PDE} for a detailed argument. It follows easily from the definition that $Z^{a,b}_{r,t}$ is additive and bounded for all $f$ in ${\cal C}^{\infty}_c(\R^n)$. \hfill $\square $\\\ For diffusions with Levy generators we introduce now another hypothesis. \begin{definition} Hypothesis $M_c$. The map $M$ with values in $\sigma$ finite measures on $\R^n-\{0\}$ satisfies hypothesis $M_C$ if it satisfies hypothesis $M$ and if \begin{equation} \sup_{s,x}\int [||y||^2 1_{||y|| \leq 1}+||y||1_{||y|| >1}]M(s,x,dy) \leq C \label{HMc} \end{equation} \label{defMc} \end{definition} The following Lemma is proved in \cite{LM} (cf the proof of Theorem 20). \begin{lemma} Let $A,B,C,D$ be strictly positive. For all $\epsilon>0$ there is $K>0$ and for all $\epsilon,\eta>0$, there is $h>0$ such that for all $0 \leq r \leq T$, for all $a$ continuous strictly elliptic bounded by $A$ for all $b$ continuous bounded by $B$ and $M$ satisfying hypothesis $M_C$, and $||y|| \leq D$, \begin{equation} Q^{a,b,M}_{r,y}[\sup_{r \leq v\leq T}||X_v||> K] \leq \epsilon \label{eqLM0} \end{equation} \begin{equation} \forall u \in [r,T],\;\; Q^{a,b,M}_{r,y}[\sup_{r \leq u \leq v \leq inf(u+h,T)}||X_v-X_u||> \eta] \leq \epsilon \label{eqLM} \end{equation} \label{lemmaLM} \end{lemma} Notice that for given $r$ and $y$, the above Lemma implies the relative weak compacity of the set of probability measures $Q^{a,b,M}_{r,y}$ satisfying the above conditions. The equation (\ref{eqLM}) is a ``uniform right convergence''. This property is stronger than the property needed to prove the relative weak compacity (Theorem 13.2 of \cite{Bil}). In particular it allows to prove the following proposition.\\ If $s_n>s$, the probability measure $Q^{a,b,M}_{s_n,y_n}$ on $\Omega^{s_n}$ can be identified with a probability measure on $\Omega^s$ still denoted $Q^{a,b,M}_{s_n,y_n}$ such that $Q^{a,b,M}_{s_n,y_n}(\{X_u=y_n,\;\;\forall s \leq u \leq s_n\})=1$. \begin{notation} Let $0 \leq r \leq s\leq t \leq T$. \begin{itemize} \item Denote $\pi^r_{[s,t]}$ the canonical projection of $\Omega^r=\Omega^r_T$ onto $\Omega^s_t$. \begin{equation} \pi^r_{[s,t]}(\omega)=\omega_{|[s,t]} \label{eqpirs} \end{equation} \item In case $r=s$ the projection $\pi^r_{[s,t]}$ will be denoted simply $\pi^r_t$ \end{itemize} \label{nota0} \end{notation} \begin{proposition} Let $a$ be continuous strictly elliptic bounded by $A$, $b$ continuous bounded by $B$ and $M$ satisfying hypothesis $M_C$. Assume that $(s_n,y_n)$ has the limit $(s,y)$. Then \begin{enumerate} \item - If $s_n$ is decreasing to $s$, with the above identification, $Q^{a,b,M}_{s_n,y_n}$ converges weakly to $Q^{a,b,M}_{s,y}$. \\ - If $s_n$ is increasing to $s$, the image of $Q^{a,b,M}_{s_n,y_n}$ by $\pi^{s_n}_{[s,T]}$ converges weakly to $Q^{a,b,M}_{s,y}$. \item For all $(t_1,..., t_k)$, $s \leq t_1<...<t_k \leq T$, for all $f$ continuous bounded on $(\R^n)^k$, $Q^{a,b,M}_{s_n,y_n}(f(X_{t_1},...,X_{t_k}))$ has the limit $Q^{a,b,M}_{s,y}(f(X_{t_1},...,X_{t_k}))$. \end{enumerate} \label{lemmFell} \end{proposition} {\bf Proof} From Lemma \ref{lemmaLM} and from Theorem 13.2 of \cite{Bil}), it follows that\\ $\{Q^{a,b,M}_{s_n,y_n},\; n \in \N\}$ in case $s_n \geq s$, ($\{(Q^{a,b,M}_{s_n,y_n})(\pi^{s_n}_{[s,T]})^{-1},\; n \in \N\}$) in case $s_n < s$ is relatively compact for the weak topology. The set of probability measures on $(\Omega^s,{\cal B}^s_T)$ is metrizable for the weak topology, thus there is a subsequence $Q_k$ converging to $Q$ for the weak topology on $\Omega^{s}$. As in Section 13 of \cite{Bil} denote $T_Q$ the set of $t \in [s,T]$ for which the projection $\Pi_t:\omega \in \Omega^s \rightarrow \omega(t)$ is continuous except at points forming a $Q$-null set. It is proved in Section 13 of \cite{Bil} that $s,T$ belong to $T_Q$ and that the complement of the set $T_Q$ in $[s,T]$ is at most countable. \begin{enumerate} \item Step 1. We prove that equations (\ref{eqLM0}) and (\ref{eqLM}) are also satisfied by $Q$ for $r=s$. Notice that the weak convergence of $Q_k$ to $Q$ means that $Q_k(f)$ has the limit $Q(f)$ for continuous functions $f$ but this convergence is not valid for general Borelian functions. Let $T^i_Q$ be an increasing sequence of finite subsets of $T_Q$ containing $s$ and $T$ such that $\cup_i T^i_Q$ is dense in $[s,T]$. It follows from the Mapping Theorem (Theorem 2.7. of \cite{Bil}) and the inequality $R(G) \leq \liminf R_k(G)$ for all open set $G$ and every sequence $R_k$ weakly converging to $R$, that for all i, \begin{equation} Q[\sup_{s\leq v \leq T,\; v \in T^i_Q}||X_v||> K] \leq \epsilon \label{eqLM2} \end{equation} \begin{equation} \forall u \in [s,T] \cap T^i_Q,\;\;Q[\sup_{u \leq v \leq (inf(u+h,T),\; v \in T^i_Q}||X_v-X_u||> \eta] \leq \epsilon \label{eqLM3} \end{equation} It follows from the monotone convergence Theorem that one can replace in the above equations $T^i_Q$ by $\cup_i T^i_Q$. Equations (\ref{eqLM0}) and (\ref{eqLM}) follow then for $Q$ making use of the density of $\cup_i T^i_Q$ in $[s,T]$ and of the right continuity of $X_v$ for all $v$. \item step 2: We prove that for all $(t_1,..., t_l)$, $s \leq t_1<...<t_l \leq T$, for all $f$ continuous bounded on $(\R^n)^l$, $Q_k(f(X_{t_1},...,X_{t_l})$ has the limit $Q(f(X_{t_1},...,X_{t_l}))$. \\ From \cite{Bil} Section 13 this is true for all $t_1,..., t_l$, in $T_Q$, $s \leq t_1<...<t_l \leq T$. The set $T_Q$ is dense and contains $s$ and $T$, thus for all $U=(t_1,..,t_l)$ there is a sequence $U^j=(t^j_1,..t^j_l)$ of $l$-uplets $T_Q$ valued decreasing to $U$. The function $f$ being uniformly continuous on compact sets, it follows from equations (\ref{eqLM0}) and (\ref{eqLM}) satisfied for all $Q_k$ and $Q$, that for all $n$, for all $j\geq J$, all $k$, $Q_k(|f(X_{t_1},...,X_{t_l})-f(X_{t^j_1},...,X_{t^j_l})|)<\epsilon$ and $Q(|f(X_{t_1},...,X_{t_l})-f(X_{t^j_1},...,X_{t^j_l})|)<\epsilon$. From the convergence of $Q_k(f(X_{t^J_1},...,X_{t^J_l})$ to $Q(f(X_{t^J_1},...,X_{t^J_l}))$, we then get the result. \item Step 3. Let $s \leq t \leq T$. Let $Z^{a,b,M}_{s',t}$ be given by equation (\ref{eqmart2}) for some function $f$ ${\cal C}^{\infty}$ with compact support. It follows from hypothesis $M_C$ that the function $(u,x) \in [0,T] \times \R^n \rightarrow L^{a,b}_u(f)(x)+K^M_u(f)(x)$ is continuous bounded. A similar argument as the above one proves that for all $\epsilon>0$, there are $s \leq s_1 < ..s_p \leq T$ and a continuous bounded function $g_{s,t}$ on $(\R^n)^p$ such that for all $Q$ and $Q_k$, $Q_k(|Z^{a,b,M}_{s,t}-g_{s,t}(X_{s_1}...X_{s_p})|\leq \epsilon$, and $Q(|Z^{a,b,M}_{s,t}-g_{s,t}(X_{s_1}...X_{s_p})|\leq \epsilon$. Making use of the martingale property of $(Z^{a,b,M}_{s,t})_{s \leq t \leq T}$ for $Q_k$ it follows that $(Z^{a,b,M}_{s,t})_{s \leq t \leq T}$ is a martingale for $Q$. \item step 4. $Q_k$ is weakly converging to $Q$ on $\Omega^{s}$. $X_s$ is continuous on $\Omega^s$, thus for all $\eta>0$, $\{||X_s-y||>\eta\}$ is open. It follows that $Q(\{||X_s-y||>\eta\}) \leq \liminf Q_k(\{||X_s-y||>\eta\})$.\\ -If $s_n<s$, $Q_k$ is the image of $Q^{a,b,M}_{s_{n_k},y_{n_k}}$. let $\epsilon$, $\eta >0$. Let $h$ such that equation (\ref{eqLM}) is satisfied for all $Q^{a,b,M}_{s_{n_k},y_{n_k}}$ and $Q$. Let $k_0$ such that for all $k \geq k_0$, $|s_{n_k}-s|<h$ and $||y-y_{n_k}||<\eta$. It follows from (\ref{eqLM}) applied with $Q^{a,b,M}_{s_{n_k},y_{n_k}}$, $u=s_{n_k}$ and $v=s$ that $Q^{a,b,M}_{s_{n_k},y_{n_k}}(\{||X_s-y||>2\eta\}) < 2 \epsilon$. $Q$ is the weak limit of $Q^{a,b,M}_{s_{n_k},y_{n_k}}(\pi^{s_n}_{[s,T]})^{-1}$, and $s$ belongs to $T_Q$, it follows that $Q(\{X_s=y\})=1$. \\ - If $s_n$ is decreasing to $s$, $Q_k=Q^{a,b,M}_{s_{n_k},y_{n_k}}$. Let $\epsilon>0$, for $k$ large enough, $Q_k(\{|X_s-y|>\epsilon\})=0$. In both cases this proves that $Q=Q^{a,b,M}_{s,y}$. \end{enumerate} \hfill $\square $\\ \begin{proposition} Let $a$ and $b$ be continuous bounded on $([0,T] \times \R^n)$ such that $a(s,x)$ is invertible for all $(s,x)$. Assume that $M$ satisfies the hypothesis $M_C$. $Z^{a,b,M}_{r,t} $ given by equation (\ref{eqmart2}) is bounded. The unique solution $Q^{a,b,M}_{r,y}$ on $({\cal D}^r, {\cal B}^r_T)$ to the additive martingale problem $Z^{a,b,M}_{r,t}$ starting from $y$ at time $r$ is weakly continuous. The set \begin{eqnarray} \Theta=\{(a,b,M) \;a,b\;\text{continuous bounded,}\;a \;\text{invertible},\nonumber\\\;M\;\text{ satisfies hypothesis $M_C$ for some $C>0$} \} \label{eqThetad} \end{eqnarray} satisfies hypothesis $H_{\Theta}$. \label{propFellL} \end{proposition} {\bf Proof} The existence and unicity follow from \cite{St}. The weak continuity follows then from Proposition \ref{lemmFell} applied with $s_n=r$ for all $n$. As already noticed in step 3 of Proposition \ref{lemmFell}, for all $f \in {\cal C}^{\infty}_c$, and all $s \leq t$, $Z^{a,b,M}_{s,t}$ is bounded. \hfill $\square $.\\ Ths weak continuity for given $r$ can also be found in \cite{EK2} Lemma 5.2., but there is no detailed proof. \\ \section{Canonical representation of the conditional expectation for the weakly continuous solution to a martingale problem} \label{seccano} Let $0 \leq r <s \leq T$. The state space $\Omega^r_T$ is either ${\cal C}([r,T],\R^n)$ or ${\cal D}([r,T],\R^n)$. The goal of this Section is to prove that in the setting of weakly continuous solutions to martingale problems, there exists a canonical choice for a regular conditional probability, and that this choice satisfies nice properties.\\ \subsection{Unicity of the solution to a martingale problem for non Markovian parameters} Let $\Theta$ be a set of parameters satisfying Hypothesis $H_{\Theta}$ (Definition \ref{not01}). Given $r \in [0,T]$ and $y \in \R^n$, we introduce the following stable set of parameters: \begin{definition} Let $0 \leq r \leq T$. ${\Theta}^r_T$ denotes the set of ${\cal B}^r_T$ predictable processes $\gamma$ on $(\Omega^r,{\cal B}^r_T)$ such that:\\ There is a finite subdivision $r=s_0<s_1<...<s_n=T$.\\ For all $i \in \{0,1,...n-1\}$ there is a finite set $I_i$, $I_0=\{1\}$, a finite partition $(A_{i,j})_{j \in I_i}$ of $\Omega^r$ into ${\cal B}^r_{s_i}$-measurable sets, and $\theta_0,\;(\theta_{i,j})_{j \in I_i} \in \Theta$ such that \begin{equation} \forall s_i \leq u < s_{i+1}, \;\forall \omega \in \Omega^r,\;\gamma(u,\omega)=\sum_{j \in I_i} \theta_{i,j}(u,X_u(\omega))1_{A_{i,j}}(\omega) \label{eqstable} \end{equation} \label{Slambda} \end{definition} We define the associated martingale: \begin{definition} For all $\gamma$ in ${\Theta}^r_T$ admitting the above representation, denote $(Z^{\gamma}_{r,u})_{r \leq u \leq T}$ the additive process $(Z^{\gamma}_{r,v}=Z^{\gamma}_{r,u}+Z^{\gamma}_{u,v}$ for $r \leq u \leq v \leq T$) defined by \begin{equation} \forall s_i \leq u \leq s_{i+1}, \;\forall \omega \in \Omega,\;Z^{\gamma}_{s_i,u}(\omega)=\sum_{j \in I_i} Z^{\theta_{i,j}}_{s_i,u}(\omega)1_{A_{i,j}}(\omega) \label{eqstable2} \end{equation} \label{defZgamma} \end{definition} \begin{lemma} Assume that $Q^{\gamma}_{r,x}$ is a solution to the bounded additive martingale problem $(Z^{\gamma}_{r,t})_{r \leq t \leq T}$ starting from $x$ at time $r$ (Definition \ref{defFellprop}). Let $(Q^{\gamma}_{r,x})^{s,\omega}$ be a regular conditional distribution of $Q^{\gamma}_{r,x}$ given ${\cal B}^r_s$. There is subset $N$ of $\Omega^r_T$ such that for all $\omega \in \Omega^r_T-N$, $(Q^{\gamma}_{r,x})^{s,\omega}$ is solution on $\Omega^r_T$ to the martingale problem $(Z^{\gamma}_{s,t})_{s \leq t \leq T}$ and $(Q^{\gamma}_{r,x})^{s,\omega}(\{X_u(\omega')=X_u(\omega),\; \forall r \leq u \leq s\})=1$. \label{lemmaunic} \end{lemma} {\bf Proof} Let $s \leq u \leq v \leq t$. For all $B \in {\cal B}^r_u$ , there is a set $N_{B}$ such that for $Q^{\gamma}_{r,x}(N_{B})=0$ and for all $\omega \in \Omega^r_T - N_{B}$, \begin{eqnarray} (Q^{\gamma}_{r,x})^{s,\omega}(Z^{\gamma}_{s,v}-Z^{\gamma}_{s,u})1_B)\nonumber\\=E_{Q^{\gamma}_{r,x}}((Z^{\gamma}_{r,v}-Z^{\gamma}_{r,u})1_B|{\cal B}^r_s)) \nonumber\\ E_{Q^{\gamma}_{r,x}}(E_{Q^{\gamma}_{r,x}}((Z^{\gamma}_{r,v}-Z^{\gamma}_{r,u})1_B|{\cal B}^r_u)|{\cal B}^r_s)=0 \label{eqrc} \end{eqnarray} The $\sigma$-algebra ${\cal B}^r_u$ is a Borel $\sigma$-algebra. Thus it is countably generated. Let $A_k$ be a countable family of events including $\Omega^r_T$ generating ${\cal B}^r_u$. It follows that there is a subset $N$ of $\Omega^r_T$ such that $Q^{\gamma}_{r,x}(N)=0$ and \begin{equation} (Q^{\gamma}_{r,x})^{s,\omega}(Z^{\gamma}_{s,v}-Z^{\gamma}_{s,u})\xi)=0 \label{eqmartin} \end{equation} for all $\xi=1_{A_k}$ and all $\omega \in \Omega^r_T-N$. It follows from the dominated convergence Theorem and the monotone class theorem that equation (\ref{eqmartin}) is satisfied by $1_A$ for all $A$ in ${\cal B}^r_u$. Thus equation (\ref{eqmartin}) is satisfied for all $\xi$ simple ${\cal B}^r_u$ measurable. Another application of the dominated convergence Theorem proves that it is satisfied for all $\xi$ bounded ${\cal B}^r_u$ measurable. We have thus proved that for all $s \leq u \leq v \leq t$, there is a set $N_{u,v}$, $Q^{\gamma}_{r,x}(N_{u,v})=0$ such that equation (\ref{eqmartin}) is satisfied for all $\xi$ bounded ${\cal B}^r_u$ measurable and all $\omega \in \Omega^r_T-N_{u,v}$. Let $N=\cup_{u,v \in ({\cal Q}\cap[s,T]) \cup \{s,T\}} N_{u,v}$. Making use of the rightcontinuity of the process $(Z^{\gamma}_{r,u})_{r \leq u \leq t}$ for given $r$, equation (\ref{eqmartin}) follows for all $s \leq u \leq v \leq t$, for all $\xi$ bounded ${\cal B}^r_u$ measurable and all $\omega \in \Omega^r_T-N$. Thus for all $\omega \in \Omega^r_T-N$(, $Q^{\gamma}_{r,x})^{s,\omega}$ is solution on $\Omega^r_T$ to the martingale problem $(Z^{\gamma}_{s,t})_{s \leq t \leq T}$. By definition of a regular conditional distribution, $(Q^{\gamma}_{r,x})^{s,\omega}(\{X_u(\omega')=X_u(\omega),\; \forall r \leq u \leq s\})=1$.\hfill $\square $ \begin{proposition} For all process $\gamma$ in $\Theta^r_T$, for all $0 \leq r<t \leq T$ there is at most one probability measure on $(\Omega^r_t,{\cal B}^r_t)$ solution to the additive martingale problem $(Z^{\gamma}_{r,u})$ starting from $y$ at time $r$. \label{propuni} \end{proposition} {\bf Proof} Let $\gamma$ in $\Theta^r_T$ (Definition \ref{Slambda}). The proof is done by iteration on $n \geq 1$. For $n=1$ it is satisfied by hypothesis on $\Theta$. Assume that it is satisfied for $n-1$. Let $Q$ be a probability measure solution to the martingale problem $(Z^{\gamma}_{r,u})$ on $(\Omega^r_t,{\cal B}^r_t)$ starting from $y$ at time $r$. It is easy to see that the restriction of $Q$ to ${\cal B}^r_{s_{n-1}}$ is solution to the martingale problem $(Z^{\tilde \gamma}_{r,u})_{r \leq u \leq s_{n-1}}$ where $\tilde \gamma$ is the restriction of $\gamma$ to $[r,s_{n-1}[ \times \Omega^r_t$. By recursion hypothesis, the restriction of $Q$ to ${\cal B}^r_{s_{n-1}}$ is uniquely determined. Let $Q^{s_{n-1},\omega}$ be a regular conditional distribution of $Q$ given ${\cal B}^r_{s_{n-1}}$. From Lemma \ref{lemmaunic}, for all $j \in I_{n-1}$, there is a subset $N_j$ of $A_{n-1,j}$ such that for all $\omega \in A_{n-1,j}-N_j$, $Q^{s_{n-1},\omega}$ is solution to the martingale problem $(Z^{\theta_{n-1,j}}_{s_{n-1},v})_{s_{n-1} \leq v \leq t}$. The function $\theta_{n-1,j}$ belongs to $\Theta$, thus from unicity of the solution to the martingale problem associated to $\theta_{n-1,j}$ starting from $X_{s_{n-1}}(\omega)$ at time $s_{n-1}$, it follows that $Q_{s_{n-1},\omega}$ is uniquely determined for all $\omega \in A_{n-1,j}-N_j$ with $Q(N_j)=0$, for all $j \in I_{n-1}$. Thus $Q$ is uniquely determined. Therefore there exists at most one probability measure $Q$ solution to the martingale problem $(Z^{\gamma}_{r,u})$ on $(\Omega^r_t,{\cal B}^r_t)$ starting from $y$ at time $r$. \hfill $\square $ \subsection{Construction of a probability measure $\overline Q$ on $(\Omega^r,{\cal B}^r_T)$ associated to a non Markovian parameter} \label{secovQ} For every Borelian function $f$ on $\Omega^r_s \times \Omega^s_T$, for all $\omega$ in $\Omega^r_s$, $f^{\omega}$ denotes the Borelian map defined on $\Omega^s_T$ by $f^{\omega}(\omega')=f(\omega,\omega')$. \begin{lemma} For given $s$, let $(Q_{s,y})_{y \in \R^n}$ be the weakly continuous solution on $(\Omega^s,{\cal B}^s_T$) to a martingale problem starting from $y$ at time $s$ (Definition \ref{defFellprop}). Then for all $f$ Borelian bounded on $\Omega^r_s \times \Omega^s_T$, the map \begin{eqnarray} \Omega^r_s \times \R^n &\rightarrow& \R \nonumber\\ (\omega,y) &\rightarrow&Q_{s,y}(f ^{\omega}) \label{eqLC2} \end{eqnarray} is measurable when $\Omega^r_s \times \R^n$ is endowed with the Borel product $\sigma$-algebra.\\ When $f$ is continuous on $\Omega^r_s \times \Omega^s_T$, the above map is furthermore separately continuous in each variable. \label{lemmaCCR2} \end{lemma} {\bf Proof} \begin{itemize} \item Assume that $f$ is continuous bounded. Being $f^{\omega}$ continuous, the map \begin{eqnarray} y\rightarrow Q_{s,y}(f^{\omega}) \;\; \label{contFell} \end{eqnarray} is continuous by hypothesis. On the other hand, $Q_{s,y}$ being a probability measure and $f$ being continuous bounded, the continuity of the map \begin{eqnarray} \omega \rightarrow Q_{s,y}(f^{\omega}). \label{contFellb} \end{eqnarray} follows from the dominated convergence Theorem. Thus the map \begin{equation} (\omega,y) \rightarrow Q_{s,y}(f^{\omega}) \label{eqLC2aa} \end{equation} is separately continuous in each variable. The sets $\Omega^r_s$ and $\Omega^s_T$ are metrizable and separable, it follows from Lemma 4.50 of \cite{AB} that this map is jointly measurable for the product Borel $\sigma$-algebra. \item Let ${\cal H}$ be the set of bounded Borelian functions $f$ on $\Omega^r_s \times \Omega^s_T$ such that the map $(\omega,y) \rightarrow Q_{s,y}(f^{\omega})$ is Borelian. ${\cal H}$ is a vector space containing the constant functions. If $f_n$ is an increasing sequence of non negative functions in ${\cal C}$ with limit $f$ bounded, it follows from the monotone convergence theorem that $Q_{s,y}(f^{\omega})=\lim_{n \rightarrow \infty} Q_{s,y}(f_n^{\omega})$. It follows that $f$ belongs to ${\cal H}$. From the first step of the proof ${\cal H}$ contains the class ${\cal C}$ of continuous bounded functions. The class ${\cal C}$ is stable by pointwise multiplication. It follows from the monotone class theorem as stated in \cite{RY} chapter 0 Theorem 2.2, that ${\cal H}$ contains all the bounded Borelian functions on $\Omega^r_s \times \Omega^s_T$. \hfill $\square $ \end{itemize} \begin{corollary} For all $f$ Borelian bounded on $\Omega^r_s \times \Omega^s_T$, the map \begin{equation} \Omega^r_s \rightarrow \R \label{eq} \end{equation} \begin{equation} \omega \rightarrow Q_{s,X_s(\omega)}(f^{\omega}) \label{eqLC2c} \end{equation} is Borelian. \label{corbor} \end{corollary} {\bf Proof} The map $\omega \in \Omega^r_s \rightarrow X_s(\omega) \in \R^n$ is Borelian, thus the corollary follows from Lemma \ref{lemmaCCR2}. \hfill $\square $ \begin{proposition} Let $0 \leq r <s$. Let $J$ be a finite subset of $\N$. For given $j \in J$, let $(Q^{j}_{s,y})_{y \in \R^n}$ be a weakly continuous family of probability measures on $\Omega^s_T$. Let $(h_j)_{j \in J}$ be non negative Borelian maps on $\R^n$ such that $\forall y \in \R^n,\;\; \sum_{j \in J}h_j=1$. Let $Q_1$ be a probability measure on $(\Omega^{r}_s, {\cal B}^{r}_s)$.\\ There is a unique probability measure $\tilde Q$ on $(\Omega^r_s \times \Omega^s_T ,{\cal B}^r_s \times {\cal B}^s_T)$ such that for all $f$ Borelian bounded on $\Omega^r_s \times \Omega^s_T $, \begin{equation} \tilde Q(f)=\sum_{j \in J}\int_{\Omega^{r}_s}h_j(X_s(\omega)) Q^j_{s,X_s(\omega)}(f^{\omega})dQ_1(\omega) \label{eqrc00} \end{equation} \label{propproba} \end{proposition} {\bf Proof} For all $f$ Borelian bounded it follows from Corollary \ref{corbor} that the map \begin{equation} T(f):\omega \in \Omega^{r}_s \rightarrow \sum_{j \in J} h_j(X_s(\omega)) Q^j_{s,X_s(\omega)}(f^{\omega}) \label{eqT} \end{equation} is Borelian. It is also bounded. Thus $\int_{\Omega^{r}_s} \sum_{j \in J}h_j(X_s(\omega)) Q^j_{s,X_s(\omega)}(f^{\omega})dQ_1(\omega)$ is well defined. The map \begin{equation} L:\; f \rightarrow \sum_{j \in J}\int_{\Omega^{r}_s}h_j(X_s(\omega)) Q^j_{s,X_s(\omega)}(f^{\omega})dQ_1(\omega) \label{eqrc01} \end{equation} defines a non negative linear form on the vector space of continuous bounded functions on $\Omega^r_s \times \Omega^s_T$. Let $f_n$ be a decreasing sequence of continuous bounded functions on $\Omega^r_s \times \Omega^s_T$ converging to $0$. From the dominated convergence Theorem, it follows that for all $j$ and $\omega$, $Q^j_{s, X_s(\omega)}(f_n^{\omega})$ is a decreasing bounded sequence with limit $0$. Applying again the dominated convergence Theorem, it follows that $L(f_n)$ is decreasing with limit $0$. From Daniell Stone Theorem it follows that there is a unique probability measure $\tilde Q$ on $(\Omega^r_s \times \Omega^s_T ,{\cal B}^r_s \times {\cal B}^s_T)$, such that for all $f$ continuous bounded, equation (\ref{eqrc00}) is satisfied. The equality (\ref{eqrc00}) follows then for all Borelian bounded function as in the end of the proof of Lemma \ref{lemmaCCR2}. \hfill $\square $ \begin{definition} Denote $\overline Q$ the probability measure on $\Omega^r_T$ deduced from $\tilde Q$ by the Borelian map \begin{eqnarray} i:\Omega^r_s \times \Omega^s_T& \rightarrow &\Omega^r_T \nonumber\\ (\omega, \omega')& \rightarrow & i(\omega, \omega') \label{eqtansfer1} \end{eqnarray} where \begin{eqnarray} i(\omega, \omega')(u)&=&\omega(u),\; \text{for}\; r \leq u \leq s\nonumber\\ &=&\omega'(u)+\omega(s)-\omega'(s),\;\text{for}\; s \leq u \leq T \label{eqtansfer2} \end{eqnarray} \label{deftransfer} \end{definition} \subsection{Canonical regular conditional probability for $\overline Q$ given ${\cal B}^r_s$} Recall that $\Pi^r_s$ denotes the canonical projection from $\Omega^r_T$ to $\Omega^r_s$. (cf Notation \ref{eqpirs}). \begin{proposition} \begin{itemize} \item For all $\psi$ Borelian bounded on $\Omega ^r_T$, for all $\omega$ in $\Omega^r_T$, let \begin{equation} T_s(\psi)(\omega)=\sum_{j \in J}h_j(X_s(\omega))(Q^j)_{s,X_s(\omega)}((\psi \circ i)^{\pi^r_s(\omega)}) \label{eqTQ} \end{equation} The map $T_s(\psi)$ defined on $\Omega ^r_T$ is ${\cal B}^r_s$ measurable. \item For all $\tilde Q$, and all $\psi$ Borelian bounded, $T_s(\psi)$ is a ${\cal B}^r_s$ measurable version of the $\overline Q$ conditional expectation of $\psi$ given ${\cal B}^r_s$. \\ For all $g$ ${\cal B}^r_s$ measurable, $T_s(g)=g$. \end{itemize} \label{propL2} \end{proposition} {\bf Proof} \begin{itemize} \item The map $i$ is measurable for the Borel $\sigma$-algebras on both side, it follows from Lemma \ref{lemmaCCR2} that for every Borelian map $\psi$ on $\Omega^r_T$, the map \begin{eqnarray} \Omega^r_s \times \R^n &\rightarrow& \R \nonumber\\ (\omega,y) &\rightarrow&Q^j_{s,y}((\psi \circ i)^{\omega}) \label{eqLC2b00} \end{eqnarray} is Borelian for all $j$. Notice that the map $\pi^r_s$ is measurable from $\Omega ^r_T$ endowed with the $\sigma$-algebra ${\cal B}^r_s$ into $\Omega ^r_s$ endowed with its Borel $\sigma$-algebra. Composing the above map (\ref{eqLC2b00}) with \begin{eqnarray} \Omega^r_T \times \R^n &\rightarrow& \Omega^r_s \times \R^n \nonumber\\ (\omega,y) &\rightarrow& (\pi^r_s (\omega),y) \label{eqLC2bb} \end{eqnarray} it follows that \begin{eqnarray} \Omega^r_T \times \R^n &\rightarrow& \R \nonumber\\ (\omega,y) &\rightarrow&Q^j_{s,y}((\psi \circ i)^{\pi^r_s(\omega)}) \label{eqLC2bbb} \end{eqnarray} is measurable for the $\sigma$-algebra product of ${\cal B}^r_s$ and of ${\cal B}(\R^n)$. $h_j$ a Borelian map on $\R^n$ and $X_s$ is ${\cal B}^r_s$ measurable, it follows that $T_s(\psi)$ is ${\cal B}^r_s$ measurable for all Borelian map $\psi$. \item Notice that every ${\cal B}^r_s$ measurable function $g$ defined on $\Omega^r_T$ can be factorized $g=\tilde g\circ\pi^r_s$ for some $\tilde g$ ${\cal B}^r_s$ measurable defined on $\Omega^r_s$. This result is deduced from the monotone class theorem (\cite{RY}) applied with the class ${\cal C}$ of functions $f(X_{t_1},X_{t_2},..,X_{t_k})$ ,$k \in \N^*$, $f$ Borelian on $(\R^n)^k$, $r \leq t_1 \leq t_2 \leq t_k \leq s$. For all $\omega$ in $\Omega^r_T$ and $\omega'$ in $\Omega^s_T$, $\pi^r_s \circ i(\pi^r_s(\omega),\omega')=\pi^r_s(\omega)$. It follows that for all $g$ ${\cal B}^r_s$ measurable, the map $(g\circ i)^{\pi^r_s(\omega)}$ is constant equal to $g(\omega)$. From the equality $\sum_{j \in J}h_j=1$, it follows that for all $g$ ${\cal B}^r_s$ measurable, $T_s(g)=g$.\\ $T_s(\psi)$ factorizes, $T_s(\psi)=\tilde T_s(\psi) \circ \pi^r_s$ where $\tilde T_s(\psi)$ is defined on $\Omega^r_s$ by \begin{equation} \tilde T_s(\psi)(\omega)= \sum_{j \in J}h_j(X_s(\omega)) Q^j_{s,X_s(\omega)}(\psi \circ i)^{\omega} \label{eq002} \end{equation} From the definition of $\overline Q$ (Proposition \ref{propproba} and Definition \ref{deftransfer}), it follows that for all $\Psi$ ${\cal B}^r_T$ measurable , and $g=\tilde g \circ \pi^r_s$ ${\cal B}^r_s$ measurable defined on $\Omega^r_T$, \begin{equation} \overline Q(\psi g)=\sum_{j \in J}\int_{\Omega^{r}_s}(\tilde g)(\omega)h_j(X_s(\omega)) Q^j_{s,X_s(\omega)}(\psi \circ i)^{\omega})dQ_1(\omega) \label{eqrc001} \end{equation} From the equality $T_s(\psi)=\tilde T_s(\psi)\circ \Pi^r_s$, it follows that \begin{equation} \overline Q(T_s(\psi) g)=\int_{\Omega^{r}_s}(\tilde g)(\omega)(\tilde T_s(\psi)(\omega)dQ_1(\omega) \label{eqrc003} \end{equation} The map $T_s(\psi)$ being ${\cal B}^r_s$ measurable, it follows from equations (\ref{eqrc001}), (\ref{eq002}) and (\ref{eqrc003}) that $T_s(\psi)$ is a ${\cal B}^r_s$ measurable version of the $\overline Q$ conditional expectation of $\psi$ given ${\cal B}^r_s$. \hfill $\square $ \end{itemize} \subsection{Existence of a solution to the martingale problem for non Markovian parameters} The goal of this Section is to prove that for all $\gamma \in \Theta^r_T$ (Definition \ref{Slambda}), the martingale problem $Z^{\gamma}$ introduced in Definition \ref{defZgamma} has a solution. \begin{proposition} The hypothesis are the same as in Proposition \ref{propproba}. Assume furthermore that $Q^{1}=Q^{1}_{r,x}$ is the unique solution on $\Omega^r_s$ to the $(Z^{\theta}_t)_{r\leq t}$ martingale problem starting from $x$ at time $r$. Assume that for all $j \in J$, $Q^{j}_{s,y}$ is the unique solution on $\Omega^{s}_T$ to the $(Z^{\theta_j}_t)_{s \leq t}$ martingale problem starting from $y$ at time $s$. Assume that there is a finite partition $(A_j)_{j \in J}$ of $\Omega^r$ in ${\cal B}^r_s$ measurable sets such that for all $j \in J$, $h_j=1_{A_j}$. Assume that for all $j$ in $J$, $(Q^{j}_{s,y})_{y \in \R^n}$ is weakly continuous. The probability measure $\overline Q$ is the unique solution to the $Z^{\beta}$ martingale problem starting from $x$ at time $r$ with $\beta(u,x)=\theta(u,x),\;$ for all $r \leq u <s$ and $\beta(u,x)=\sum_{j \in J}1_{A_j}\theta_j(u,x)$,\; for all $s \leq u$. \label{propcondprob} \end{proposition} {\bf Proof} \begin{itemize} \item Let $s \leq t<u \leq T$. By definition of $\beta$ and the additive property of the martingales, $Z^{\beta}_{r,u}-Z^{\beta}_{r,t}=Z^{\beta}_{t,u}= \sum_{j \in J}1_{A_j} (Z^{\theta_j}_{s,u}-Z^{\theta_j}_{s,t})$. For all $\omega$ in $\Omega^r_s$ and $\omega'$ in $\Omega^s_T$, such that $X_s(\omega)=X_s(\omega')$, for all $s \leq v \leq T$, $i(\omega,\omega')(v)=\omega'(v)$, and $[(Z^{\theta_j}_{s,u}-Z^{\theta_j}_{s,t}) \circ i]^{\omega}(\omega')= (Z^{\theta_j}_{s,u}-Z^{\theta_j}_{s,t})(\omega')$. Notice also that for all $\psi$ ${\cal B}^r_t$ measurable and all $\omega$ in $\Omega^r_s$, the map $(\psi \circ i)^{\omega}$ is ${\cal B}^s_t$ measurable. Thus for all $\psi$ ${\cal B}^r_t$ measurable \begin{eqnarray} 1_{A_j}(X_s(\omega))Q^j_{s,X_s(\omega)}[([\psi (Z^{\beta}_{r,u}-Z^{\beta}_{r,t}]\circ i)^{\omega}]\nonumber\\ =1_{A_j}(X_s(\omega))Q^j_{s,X_s(\omega)} [1_{\{\omega'|\; X_s(\omega)=X_s(\omega')\}}(\psi \circ i)^{\omega}[(Z^{\theta_j}_{s,u}-Z^{\theta_j}_{s,t}) \circ i]^{\omega}]\nonumber\\ =1_{A_j}(X_s(\omega)Q^j_{s,X_s(\omega)}[(\psi \circ i)^{\omega}(Z^{\theta_j}_{s,u}-Z^{\theta_j}_{s,t})] \label{eqmart1} \end{eqnarray} $(Z^{\theta_j}_{s,t})_{s \leq t \leq T}$ is a $Q^j_{s,y}$ martingale for all $y$, it follows that the last term of (\ref{eqmart1}) is equal to $0$. It follows then from (\ref{eqmart1}) and the definition of $\overline Q$ (Section \ref{secovQ}) that $Z^{\beta}_{r,t}$ is a $\overline Q$ martingale for $s \leq t \leq T$. \item Let $r \leq t<u \leq s$. $Z^{\beta}_{r,u}-Z^{\beta}_{r,t}=(Z^{\theta}_{r,u}-Z^{\theta}_{r,t})=(Z^{\theta}_{r,u}-Z^{\theta}_{r,t})\circ (\pi^r_s)$. Let $\psi$ be a function ${\cal B}^r_t$ measurable on $\Omega^r_T$, $\psi=\tilde \psi \circ (\pi^r_s)$, where $\tilde \psi$ is a function ${\cal B}^r_t$ measurable defined on $\Omega^r_s$. Being $\psi$ and $Z^{\beta}_{r,u}-Z^{\beta}_{r,t}$ ${\cal B}^r_s$ measurable, it follows from the second part of the proof of Proposition \ref{propL2} that $\overline Q(\psi (Z^{\beta}_{r,u}-Z^{\beta}_{r,t}))=Q_1(\tilde \psi (Z^{\theta}_{r,u}-Z^{\theta}_{r,t}))$. Being $(Z^{\theta}_{r,t})_{r \leq t \leq s}$ a $Q_1$ martingale, it follows that $Z^{\beta}_{r,t}$ is a $\overline Q$ martingale for $r \leq t \leq s$. \hfill $\square $ \end{itemize} \subsection{Canonical regular conditional probability} \begin{notation} For all $r \leq s$, for all $\omega \in \Omega^r_T$ and $\omega' \in \Omega^s_t$ such that $X_s(\omega)=X_s(\omega')$ we denote $\omega*\omega'$ the element of $\Omega^r_T$ such that \begin{eqnarray} \omega*\omega'(u)&=&\omega(u)\;\; \forall r \leq u \leq s\nonumber\\ \omega*\omega'(u)&=&\omega'(u)\;\; \forall s <u \leq T \label{eq*} \end{eqnarray} \label{not*} \end{notation} \begin{proposition} \begin{enumerate} \item Let $r \geq 0$ and $y$ in $\R^n$. Let $\theta \in \Theta$. For all $s \in [r,T[$ and all bounded map $\psi$ ${\cal B}^r_T$-measurable defined on $\Omega^r$ , the map $T^{\theta}_s(\psi)$ defined on $\Omega^r$ by $T^{\theta}_s(\psi)(\omega)=Q^{\theta}_{s, X_s(\omega)}((\psi \circ i)^{\pi^r_s(\omega)})$ is a ${\cal B}^r_s$-measurable version of the $Q^{\theta}_{r,y}$-conditional expectation of $\psi$ given ${\cal B}^r_s$. We call it the canonical regular conditional distribution of $Q^{\theta}_{r,y}$. \item \begin{eqnarray} T^{\theta}_s(\psi)(\omega)&=&Q^{\theta}_{s, X_s(\omega)}((\psi \circ i)^{\pi^r_s(\omega)})\label{eq39a}\\&=&\int_{\Omega^s_T} \psi(\omega*\omega')1_{X_s(\omega)=X_s(\omega')}dQ^{\theta}_{s, X_s(\omega)}(\omega') \label{eqTs} \end{eqnarray} \item It satisfies the following chain rule for all map $\phi$ ${\cal B}^{r_0}_T$-measurable defined on $\Omega^{r_0}$, \begin{equation} \forall r_0\leq r \leq s \leq T,\;\; T^{\theta}_r(\phi)=T^{\theta}_r(T^{\theta}_s(\phi)) \label{eqcomp} \end{equation} \end{enumerate} \label{proCRC} \end{proposition} {\bf Proof} \begin{enumerate} \item Apply Proposition \ref{propcondprob} with $Q_1=Q^{\theta}_{r,y}$, $J=\{j\}$, and $Q^j_{s,x}= Q^{\theta}_{s,x}$. It follows that the probability measure $\overline Q$ satisfies the $Z^{\theta}$ martingale problem on $\Omega^r_T$ starting from $y$ at time $r$. By unicity of the solution to the martingale problem it follows that $\overline Q=Q^{\theta}_{r,y}$. The result follows then from Proposition \ref {propL2}. \item It follows immediately from the definitions of $i$, $\Pi^r_s$ and $*$ that \begin{equation} 1_{\{X_s(\omega)=X_s(\omega')\}}(\psi \circ i)^{\pi^r_s(\omega)}(\omega')=1_{\{X_s(\omega)=X_s(\omega')\}}\psi(\omega*\omega') \label{eq*2} \end{equation} The equation (\ref{eqTs}) follows then from $Q^{\theta}_{s, X_s(\omega)}(1_{\{X_s(\omega)=X_s(\omega')\}})=1$ \item Let $\overline Q$ be as in 1.. From the definition of $\overline Q$, Definition \ref{deftransfer}, equations (\ref{eqrc00}) and (\ref{eqTs}) it follows from the equality $\overline Q=Q^{\theta}_{r,y}$ that for all function $\psi$ ${\cal B}^{r}_T$-measurable defined on $\Omega^{r}$ , \begin{equation} \int_{\Omega^r_T} \psi (\omega')dQ^{\theta}_{r,y}(\omega')=\int_{\Omega^r_s}[ \int_{\Omega^s_T}\psi(\omega_1*\omega_2) 1_{X_s(\omega_1)=X_s(\omega_2)}dQ^{\theta}_{s, X_s(\omega_1)}(\omega_2)] dQ^{\theta}_{r,y}(\omega_1) \label{eqcr1} \end{equation} Let $\phi$ ${\cal B}^{r_0}_T$ measurable defined on $\Omega^{r_0}$. Let $\omega \in \Omega^{r_0}_r$. Applying the above equation to the function $\psi$ defined on $\Omega^r$ by $\psi(\omega')=\phi(\omega*\omega')1_{\{X_r(\omega)=X_r(\omega')\}}$, it follows from equation (\ref{eqTs}) that \begin{eqnarray} T^{\theta}_r(\phi)=\int_{\Omega^r_T} \phi (\omega*\omega')1_{X_r(\omega)=X_r(\omega')}dQ^{\theta}_{r,X_r(\omega)}(\omega')=\nonumber\\ \int_{\Omega^r_T}1_{X_r(\omega)=X_r(\omega_1*\omega_2 )}dQ^{\theta}_{r,X_r(\omega)}(\omega_1) [ \int_{\Omega^s_T}\phi(\omega*\omega_1*\omega_2)1_{X_s(\omega_1)=X_s(\omega_2)} dQ^{\theta}_{s, X_s(\omega_1)}(\omega_2)]\nonumber\\ =\int_{\Omega^r_T}1_{X_r(\omega)=X_r(\omega_1)T^{\theta}_s(\phi)(\omega*\omega_1) )}dQ^{\theta}_{r,X_r(\omega)}(\omega_1) \label{eqcr2} \end{eqnarray} This proves the chain rule equation (\ref{eqcomp}). \hfill $\square $ \end{enumerate} \begin{theorem} Let $0 \leq r \leq T$. For all $\gamma \in \Theta^r_T$ and $y$ in $\R^n$, there is a unique solution to the martingale problem $Z^{\gamma}$ on $(\Omega^r,{\cal B}^r_T)$ starting from $y$ at time $r$. We denote it $Q^{\gamma}_{r,y}$. Let $r\leq s <T$. Consider the expression of $\gamma$ as in Definition \ref{Slambda}. For all $r \leq s<T$, let $i$ such that $s_i \leq s <s_{i+1}$. For all $\psi$ defined on $\Omega^r$ bounded ${\cal B}^r_{s_{i+1}}$ measurable, the map $T^{\gamma}_s(\psi)$ defined as \begin{equation} T^{\gamma}_s(\psi)=\sum_{j \in J_i}1_{A_{i,j}}T^{\theta_{i,j}}_s(\psi) \label{Tgamma} \end{equation} is $ {\cal B}^r_{s_i}$ measurable. The map $T^{\gamma}_s$ admits a unique extension to all bounded ${\cal B}^r_T$ measurable functions such that the chain rule is satisfied: \begin{equation} \forall r \leq s \leq t \leq T,\; T^{\gamma}_s(\psi)=T^{\gamma}_s (T^{\gamma}_t(\psi)) \label{eqcomp3a} \end{equation} $T^{\gamma}_s(\psi)$ is a ${\cal B}^r_s$ measurable version of the $Q^{\gamma}_{r,y}$ conditional expectation of $\psi$ given ${\cal B}^r_s$. It is called the canonical regular conditional distribution of $Q^{\gamma}_{r,y}$ given ${\cal B}^r_s$. \label{propcanreg} \end{theorem} {\bf Proof} For given $r$ and $y$ the existence of $Q^{\gamma}_{r,y}$ follows by induction from Proposition \ref{propcondprob}. The unicity follows from Proposition \ref{propuni}. Let $s_i$ be as subdivision asociated to $\gamma$ as in Definition \ref{Slambda}. Let $ s_i \leq s <s_{i+1}\leq T$. Consider the probability measures $\overline Q$ constructed on $\Omega^r_{s_{i+1}}$ from $Q^{\gamma}_{r,y}$ on $\Omega^r_s$ and from $\sum 1_{A_{i,j}}(\omega)Q^{\theta_{i,j}}_{s,X_s(\omega)}$ on $\Omega^s_T$ as in Section \ref{secovQ}. From Proposition \ref{propcondprob} it follows that $\overline Q$ solves the martingale problem for $\gamma$ on $\Omega^r_{s_{i+1}}$ starting from $y$ at time $r$. From the unicity it follows that the restriction of $Q^{\gamma}_{r,y}$ to ${\cal B}^r_{s_{i+1}}$ is equal to $\overline Q$. From Proposition \ref{propL2} equation (\ref {eqTQ}) and equation (\ref{eq39a}), it follows that for all $\psi$ ${\cal B}^r_{s_{i+1}}$ measurable, $T^{\gamma}_s(\psi)$ defined by equation (\ref{Tgamma}) is a ${\cal B}^r_s$ measurable version of the $\overline Q$- conditional expectation of $\psi$ given ${\cal B}^r_s$. It is thus also a ${\cal B}^r_s$ measurable version of the $Q^{\gamma}_{r,y}$-conditional expectation of $\psi$ given ${\cal B}^r_s$.\\ Making use of the chain rule for $T^{\theta}$, Proposition \ref{proCRC}, it follows that $T^{\gamma}_s$ admits a unique extension to all ${\cal B}^r_T$ measurable functions such that the chain rule (\ref{eqcomp3a}) is satisfied. It is given by a stepwise evaluation: for $s_i \leq s < s_{i+1}$, $T^{\gamma}_s(\psi)=T^{\gamma}_s (T^{\gamma}_{s_{i+1}}(...T^{\gamma}_{s_{n-1}}(\psi))$. The result follows then by induction. \hfill $\square $ \begin{corollary} \begin{enumerate} \item Let $r\geq 0$, for all $\gamma \in \Theta^r_T$, and $r \leq s <T$, the non negative linear map $T^{\gamma}_s$ is continuous from below. It can be uniquely extended to the Borelian functions bounded from below by the formula \begin{equation} T^{\gamma}_s(\psi)=lim_{n \rightarrow \infty}T^{\gamma}_s(\psi \wedge n) \label{eqext} \end{equation} This extension is continuous from below. \item Equation (\ref{eqTs}) (resp (\ref{Tgamma})) is still satisfied for random variables bounded from below, for all $\theta \in \Theta$ (resp $\gamma \in \Theta^r_T$). \item $T^{\gamma}_s(\psi)$ is a ${\cal B}^r_s$-measurable version of the $Q^{\gamma}_{r,y}$-conditional expectation of $\psi$ given ${\cal B}^r_s$. \item The chain rule is satisfied for all function $\psi$ ${\cal B}^r_T$-measurable bounded from below, \begin{equation} \forall r \leq s \leq t \leq T,\;\; T^{\gamma}_s(\psi)=T^{\gamma}_s(T^{\gamma}_t(\psi)) \label{eqcompe} \end{equation} \end{enumerate} \label{corext} \end{corollary} {\bf Proof} The continuity from below for $T^{\theta}_s$, $\theta \in \Theta$ follows from equation (\ref{eqTs}) and from the monotone convergence theorem applied to the probability measure $Q^{\theta}_{s, X_s(\omega)}$ for given $\omega$. It follows for general $T^{\gamma}_s$ from the formula (\ref{Tgamma}). The other statements follow from the continuity from below of $T^{\gamma}_s$ and from the monotone convergence theorem for probability measures (or for conditional expectations to prove {\it 3.}). \hfill $\square $ We are now able to give a definition of a stable set of probability measures for probability measures which are non dominated by some probability measure. \section{Stable set of probability measures and penalties} \label{secstable} We concentrate now on the non dominated framework. As already noticed the usual definition of a stable set of probability measures cannot be used because the $Q$-conditional expectation is defined up to $Q$-null set. The existence of a canonical choice for the regular conditional probability is crucial in the construction of dynamic procedures in non dominated setting. \subsection{Stable set of probability measures} \label{susecsta} Let $0 \leq r \leq T$. Let $\Omega^r$ be ${\cal C}^r$ or ${\cal D}^r$ with the canonical filtration. For all $r \leq s\leq T$, denote $({\cal B}_b)^r_s$ the set of bounded functions defined on $\Omega^r$, ${\cal B}^r_s$-measurable. Let ${\cal T}$ be a subset of $[r,T]$. \begin{definition} A set ${\cal Q}$ of probability measures on $(\Omega^r,{\cal B}^r_T)$ is ${\cal T}$ stable if \begin{enumerate} \item Special choice of a regular conditional distribution\\ For all $Q \in {\cal Q}$ for all $r\leq s \leq T$ $s \in {\cal T}$, there is a non negative linear map continuous from below $T^Q_s: ({\cal B}_b)^r_T \rightarrow ({\cal B}_b)^r_s$ such that for all $\psi \in ({\cal B}_b)^r_T$, $T^Q_s(\psi)$ is a version of the $Q$-conditional expectation of $\psi$ given $({\cal B}_b)^r_s$. \item Chain rule\\ For all map $\psi$ in $({\cal B}_b)^r_T$, for all $s,t$ in ${\cal T}$, $s \leq t$ \begin{equation} \;\; T^{Q}_s(\psi)=T^{Q}_s(T^{Q}_t(\psi)) \label{eqcomp3} \end{equation} \item Stability by composition\\ For all $Q$ and $R$ in ${\cal Q}$, for all $s \in {\cal T}$ there is a probability measure $S$ in ${\cal Q}$ such that \begin{eqnarray} \forall s \leq u \leq T, \; T^S_u= T^R_u \nonumber\\ \forall r \leq u<s, \forall \phi \in ({\cal B}_b)^r_s,\; T^S_u(\phi)= T^Q_u(\phi) \label{eqcompo} \end{eqnarray} \item stability by bifurcation:\\ for all $ s \in {\cal T}$ for all $Q$ and $R$ in ${\cal Q}$, for all $A \subset \Omega^r$, $A \in {\cal B}^r_s$, there is a probability measure $S$ in ${\cal Q}$ such that \begin{equation} \forall s \leq u \leq T, \;T^{S}_u=1_AT^{Q}_u+1_{A^c}T^{R}_u \label{eqbif} \end{equation} \end{enumerate} \label{defstable} \end{definition} In all the following to simplify the notations we assume that ${\cal T}=[r,T]$. \begin{theorem} Assume that $\Theta$ satisfies hypothesis $H_{\Theta}$ (Definition \ref{not01}). Let $0 \leq r \leq T$ and $y$ in $\R^n$. Let $\Theta^r_T$ be the set introduced in Definition \ref{Slambda}. The set \begin{equation} {\cal Q}^{\Theta}_{r,y}=\{Q^{\gamma}_{r,y},\gamma \in \Theta^r_T\} \label{eqsta} \end{equation} is a stable set of probability measures. \\ More precisely, let $Q=Q^{\gamma}_{r,y}$ and $R=Q^{\delta}_{r,y}$ in ${\cal Q}^{\Theta}_{r,y}$. \begin* \item{i)} Let $\lambda(u,\omega)=\gamma(u,\omega)$ for all $r \leq u <s$ and $\lambda(u,\omega)=\delta(u,\omega)$ for all $s \leq u < T $. The probability measure $S=Q^{\lambda}_{r,y}$ satisfies (\ref{eqcompo}). \item{ii)} Let $\eta(u,\omega)=\gamma(u,\omega)$ for all $r \leq u < s$ and $\eta(u,\omega)=1_A(\omega)\gamma(u,\omega)+1_{A^c}\delta(u,\omega)$ for all $s \leq u < T $. The probability measure $S=Q^{\eta}_{r,y}$ satisfies (\ref{eqbif}). \end* \label{thmsta} \end{theorem} {\bf Proof} For $\gamma \in \Theta^r_T$ and $r \leq s < T$, let $T^{Q^{\gamma}_{r,y}}_s=T^{\gamma}_s$. It follows from Corollary \ref{corext} that $T^{Q^{\gamma}_{r,y}}_s$ is linear non negative and continuous from below. Properties {\it 1.} and {\it 2.} follow from Theorem \ref{propcanreg}. \\ - Proof of property {\it 3.} The process $\lambda$ defined in ${i)}$ belongs to $\Theta^r_T$ and from Theorem \ref{propcanreg} there is a unique probability measure $Q^{\lambda}_{r,y}$ solution on $(\Omega^r,{\cal B}^r_T)$ to the martingale problem associated to $\lambda$ starting from $y$ at time $r$. It follows also from equation (\ref{Tgamma}) and the chain rule that for all $r \leq u \leq v \leq T$ for all $\psi$ defined on $\Omega^r$ ${\cal B}^r_v$ measurable, the ${\cal B}^r_u$ measurable map $T^{\lambda}_u(\psi)$ depends only on the restriction of $\lambda$ to $[u,v[$. This end the proof of the stability by composition.\\ - Property {\it 4.} is proved in the same way considering the process $\eta$ defined in ${ii)}$\hfill $\square $\\ \linebreak As in \cite{JBN-PDE} we can furthermore consider a multivalued Borel mapping $\Gamma$ from $\R^+ \times\R^n$ into $E$ (definition 2.8 of \cite{JBN-PDE}). For all $(t,x)$, $\Gamma(t,x)$ is a subset of $E$ containing $0$. Let $\Theta(\Gamma)=\{\theta \in \Theta\;|\theta(t,x) \in \Gamma(t,x) \;\forall (t,x)\}$. We then get the following corollary \begin{corollary} Assume that $\Theta$ satisfies hypothesis $H_{\Theta}$. For all multivalued Borel mapping $\Gamma$, the set \begin{equation} {\cal Q}^{\Theta}(\Gamma)_{r,y}=\{Q^{\gamma}_{r,y},\gamma \in \Theta(\Gamma)^r_T\} \label{eqstab} \end{equation} is a stable set of probability measures. We call it the stable set of probability measures generated by $\Theta (\Gamma)$. \label{corsta} \end{corollary} \subsection{penalties} \label{subsecpen} We want to construct time consistent dynamic convex Feller procedures, (and not only sublinear). Therefore as for the construction of time consistent dynamic risk measures on a filtered probability space, we need to introduce penalties satisfying some properties (\cite{BN03}). To encompass the non dominated framework, we must adapt the definition of the required properties for penalties. \begin{definition} Assume that $\Theta$ satisfies hypothesis $H_{\Theta}$. Let $\Gamma$ be a multivalued Borel mapping. Let ${\cal Q}={\cal Q}^{\Theta}(\Gamma)_{r,y}$. A penalty defined on ${\cal Q}$ is a family $(\alpha_{s,t}(Q))_{r \leq s \leq t\leq T}$ of ${\cal B}^r_s$-measurable functions on $\Omega^r$ ${\cal B}^r_s$-measurable bounded from above such that \begin{enumerate} \item It is local:\\ For all $\gamma,\eta \in {\Theta(\Gamma)^r_T}$, for all $A \in {\cal B}^r_s$, if $1_A \gamma(u,\omega)=1_A \eta(u,\omega)$, for all $u \in[s,t[$, then $1_A \alpha_{s,t}(Q^{\gamma}_{r,y})=1_A \alpha_{s,t}(Q^{\eta}_{r,y})$. \item It satisfies the cocycle condition: \begin{equation} \forall Q \in {\cal Q},\;\forall r \leq s \leq t \leq u \leq T,\; \alpha_{s,u}(Q)=\alpha_{s,t}(Q)- T^Q_s(-\alpha_{t,u}(Q)) \label{eqcc} \end{equation} \end{enumerate} \label{defpen} \end{definition} \begin{remark} Penalties could be defined for more general stable sets of probability measures. \label{rempen} \end{remark} The construction that we have made in \cite{JBN-PDE} in the setting of (equivalent) probability measures solution to a martingale problem associated to a continuous diffusion leads also to penalties in the non dominated framework. This will be detailed when we construct time consistent dynamic procedures (Section \ref{secTCCC}).\\ Recall also that one of our goals is to construct time consistent convex procedures which give rise to viscosity solutions to second order partial differential equations. Therefore as in \cite{JBN-PDE} we are interested in continuity properties. \section{Continuity properties for the canonical conditional probability} \label{secconti} \subsection{A continuity property in the case of continuous paths} \label{secccp} \begin{proposition} Case of ${\cal C}$: For all $u$ in $\R^+$ and $x \in \R^n$, let $Q_{u,x}$ be the unique solution to a martingale problem. Assume that for some $s$, $(Q_{s,y})_{y \in \R^n}$ is weakly continuous. \begin{enumerate} \item Then for all $0 \leq r <s$ for all $f$ bounded uniformly continuous (for the uniform norm) on $\Omega^r_T$, the canonical conditional probability $T^{Q_{r,x}}_s(f):$ \begin{eqnarray} \Omega^r_T &\rightarrow& \R \nonumber\\ (\omega &\rightarrow&Q_{s,X_s(\omega)}((f \circ i)^{\pi^r_s(\omega}))=Q_{s,X_s(\omega)}((f(\omega*\omega')1_{X_s(\omega)=X_s(\omega')}) \label{eqCCCa} \end{eqnarray} is a continuous version of the $Q_{r,x}$ conditional expectation given ${\cal B}^r_s$. \item If $f=g \circ \pi^r_{t'}$ for some $g$ continuous on $\Omega^{r}_{t'}$ with compact support. Let $t''=inf(s,t')$. Then $Q_{s,X_s(\omega)}((f \circ i)^{\pi^r_s(\omega}))=h \circ \pi^r_{t"}$ for some continuous function $h$ on $\Omega^{r}_{t"}$ with compact support. \end{enumerate} \label{propCCCa} \end{proposition} {\bf Proof} \begin{enumerate} \item The map $i$ introduced in Definition \ref{deftransfer} is uniformly continuous for the uniform norm, it follows that for all $f$ uniformly continuous bounded on $\Omega^r_T$, the map $f \circ i$ is uniformly continuous bounded on the product space $\Omega^r_s \times \Omega^s_T$ for the uniform norm. Being $f \circ i$ uniformly continuous, for all $\epsilon>0$ there is $\eta>0$ such that for $||\omega-\omega'||<\eta$, $||(f \circ i)^{\omega}-(f \circ i)^{\omega'}||< \epsilon$. The continuity of the map \begin{eqnarray} \Omega^r_s \times \R^n &\rightarrow& \R \nonumber\\ (\omega,y) &\rightarrow&Q_{s,y}((f \circ i)^{\omega}) \label{eqLC2b} \end{eqnarray} follows then easily from the weak continuity of the family $(Q_{s,y})_{y \in \R^n}$. being $X_s$ continuous from $\Omega^r_T$ to $\R^n$ and $\pi^r_s$ continuous from $\Omega^r_T$ to $\Omega^r_s$, the continuity follows by composition. From Proposition \ref{proCRC}, equation (\ref{eqCCCa}) describes the canonical regular conditional distribution.\\ \item Assume now that $f=g \circ \pi^r_{t'}$ for some $g$ continuous on $\Omega^{r}_{t'}$ with compact support $K$. \begin{enumerate} \item If $t'\leq s$, $f$ is ${\cal B}^r_s$ measurable, and from Proposition \ref{propL2}, $ T^{Q_{r,x}}_s(f)=f$ \item If $s \leq t'$, then $t''=s$. The maps $g$ and $\pi^r_{t'}$ are uniformly continuous thus from 1., $T^{Q_{r,x}}_s(f)$ is continuous and ${\cal B}^r_s$ measurable. Therefore it can be written $T^{Q_{r,x}}_s(f)=h\circ \pi^r_s$ for some continuous function $h$. The map $\pi:\omega \in \Omega^{r}_{t'} \rightarrow \omega_{|[r,s]} \in \Omega^{r}_{s}$ is continuous. Thus $\pi(K)$ is a compact subspace of $\Omega^{r}_{s}$. If $\omega \notin \pi(K)$ and $\omega' \in \Omega^s_T$, $Q_{s,X_s(\omega)}((f(\omega * \omega')1_{ X_s(\omega)= X_s(\omega')})=0$. This proves the result with the support of $h$ contained in $\pi(K)$. \end{enumerate} \end{enumerate} \hfill $\square $ Notice that this provides a class of continuous functions stable by $T^{\theta}$: \begin{notation} In case $\Omega^r={\cal C}^r$, we denote \begin{equation} {\cal V}^r_t=\{f\circ \pi^r_u,r \leq u \leq t, f \text{continuous on $\Omega^r_u$ with compact support }\} \label{eqV} \end{equation} For all $Y \in {\cal V}^r_t$, for all $y \in \R^k$ and $s \geq r$, $T^{Q_{r,y}}_s(Y)$ belongs to ${\cal V}^r_t$ \label{notcor} \end{notation} \begin{lemma} Let $\Omega$ be a Polish space. Let ${\cal Q}$ be a set of probability measures on $(\Omega,{\cal B}(\Omega))$. Assume that ${\cal Q}$ is tight. For all $f$ continuous bounded on ${\cal C}^r$, there is an increasing sequence of continuous functions with compact support $f_n$ such that $f=\lim_{n \rightarrow \infty}f_n \; Q \; a.s. \; \forall Q \in {\cal Q}$. \label{lemmatig} \end{lemma} {\bf Proof} By hypothesis there is an increasing sequence ${\cal K}_n$ of compact subspaces of $\Omega$ such that $Q({\cal K}_n^c)< \frac{1}{n}$ for all $Q \in {\cal Q}$. Let $\phi_n$ be an increasing sequence of continuous functions with compact support, $1_{{\cal K}_n} \leq \phi_n \leq 1$. Assume that $f$ is continuous bounded, let $g=\lim_{n \rightarrow \infty} f\phi_n$. $0 \leq ||f-f\phi_n||\leq||f||$. $Q(|f-f\phi_n|)\leq||f||\frac{1}{n}$. From the dominated convergence Theorem, it follows that $f=g=\lim f \phi_n\;\; Q \; a.s. \; \forall Q \in {\cal Q}$. \subsection{Feller property} \label{secfell} \begin{definition} \begin{enumerate} \item Let $s$ in $\R^+$. For all $x$ in $\R^n$, let $Q^{\theta}_{s,x}$ be the unique weakly continuous solution to a martingale problem starting from $x$ at time $s$. \\ $(Q^{\theta}_{s,x})_{x \in \R^n}$ has the Feller property if for all $t\geq s$, for all $f$ continuous bounded on $\R^n$, the map $x \rightarrow Q^{\theta}_{s,x}(f(X_t))$ is continuous. We denote it $T_{st}(f)$. \item Assume that $\Theta$ satisfies hypothesis $H_{\theta}$. $\Theta$ has the Feller property if every $(Q^{\theta}_{s,x})$ has the Feller property. \end{enumerate} \label{defFeller} \end{definition} The sets $\Theta$ constructed in section 2 have the Feller property: \begin{lemma} The sets $\Theta$ given by equation (\ref{eqThetac}) in case of continuous diffusions and by equation (\ref{eqThetad}) in case of diffusions with Levy generator satisfy the property $H_{\theta}$ and have the Feller property. \label{lemFellprop} \end{lemma} {\bf Proof} The property $H_{\theta}$ has been proved in Proposition \ref{propFell} in case of continuous diffusions. The map $X_t$ is continuous, thus the Feller property is satisfied. In case of diffusions with Levy generator, the property $H_{\theta}$ has been proved in Proposition \ref{propFellL}. The Feller property has been proved in Proposition \ref{lemmFell}. \begin{lemma} Case of ${\cal D}$ or ${\cal C}$. let $Q^{\theta}_{s,x}$ be the unique weakly continuous solution to a martingale problem starting from $x$ at time $s$. Assume that $(Q^{\theta}_{s,x})_{x \in \R^n}$ has the Feller property. Then \begin{equation} \omega \in \Omega^r_T \rightarrow T_{st}(f)(X_s)(\omega) \label{eqDCC2} \end{equation} is the canonical version of the $Q^{\theta}_{r,x}$ conditional expectation of $f(X_t)$ given ${\cal B}^r_{s}$. \label{lemmaCC2} \end{lemma} {\bf Proof} Let $\psi=f(X_t)$, $t \geq s$. With the notations of Proposition \ref{proCRC}, $1_{X_s(\omega)=X_s(\omega')}(\psi \circ i)^{\pi^r_s(\omega)}(\omega')=f(X_t)(\omega')$. The Lemma follows then easily from Proposition \ref{proCRC}, and the definition of $T_{st}(f)$. \begin{notation} Let ${\cal Q}$ be a weakly relatively compact set of probability measures. Denote as in \cite{DHP} and \cite{BNK1} $L^1(c)$ the Banach space obtained from the completion and separation of continuous function for the c norm $c(f)=\sup_{Q \in {\cal Q}}Q(|f|)$. \label{notc} \end{notation} \begin{lemma} Let ${\cal P}$ be a set of probability measures on ${\cal C}([s,T],\R^n)$. Assume that ${\cal P}$ is tight. Let ${\cal T}$ be a dense subset of $[s,T]$. The set of continuous bounded coordinates map $f(X_{t_1},... X_{t_k})$, $s\leq t_1<...<t_k\leq T$, $t_i \in {\cal T}$ and $f \in {\cal C}_b((\R^n)^k)$ is dense in $L^1(c)$. \label{lemmacont} \end{lemma} {\bf Proof} Being ${\cal P}$ tight, there is a compact subset $K$ of $\Omega^s={\cal C}([s,T], \R^n)$ such that for all $P \in {\cal P}$, $P(\omega^s-K)< \epsilon$. Let ${\cal E}=\{f(X_{t_1},... X_{t_k}),\;s\leq t_1<...<t_k\leq t,\;t_i \in {\cal T},\; f\in {\cal C}_b((\R^n)^k)\}$. The restriction to $K$ of elements of ${\cal E}$ is an algebra which separates the points of the Haussdorf compact space $K$. Thus from Stone Weierstrass theorem it is dense in ${\cal C}(K)$. This proves the result. \begin{notation} Let $0 \leq r \leq s\leq T$. For all $j \in \N^*$ and $k \in \N^*$, for all $f$ Borelian on $(\R^n)^{j+k}$, given $r \leq s_1<...<s_j<s \leq t_1<..t_k \leq T$, denote $f_{(s_1,..,s_j),(t_1,...t_k)}$ the map defined on $\Omega^r_T$ by $f_{(s_1,..,s_j),(t_1,...t_k)}(\omega)= f(X_{s_1},..X_{s_j},X_{t_1},...X_{t_k})(\omega)$. \label{notf} \end{notation} \begin{proposition} Case of ${\cal D}$ or ${\cal C}$. For all $s$ in $\R^+$ and $x \in \R^n$, let $Q^{\theta}_{s,x}$ be the unique weakly continuous solution to a martingale problem starting from $x$ at time $s$. Assume that for all $s$ $(Q^{\theta}_{s,x})_{x \in \R^n}$ has the Feller property. Then for all $f$ continuous bounded (resp. Borelian bounded) on $(\R^n)^{j+k}$, there is a continuous bounded (resp. Borelian bounded) map $\hat f$ on $(\R^n)^{j+1}$ such that for all $r <s$ and $x$ in $\R^n$, \begin{equation} \omega \in \Omega^r_T \rightarrow \hat f((X_{s_1}(\omega),...,X_{s_j}(\omega),X_s(\omega)) \label{eqDCC2b} \end{equation} is the canonical version of the $Q_{r,x}$ conditional expectation of $f_{(s_1,..,s_j),(t_1,...t_k)}$ given ${\cal B}^r_{s}$, i.e. $T^{\theta}_s(f_{(s_1,..,s_j),(t_1,...t_k)})=\hat f_{s_1,...,s_j,s}$ \label{propDCC2} \end{proposition} {\bf Proof} We prove the proposition for all $f$ continuous bounded on $(\R^n)^{j+k}$. The result follows then for $f$ Borelian bounded from the monotone class theorem. \begin{itemize} \item Assume that $k=1$. Let $\psi=f_{(s_1,..,s_j),(t_1)}$. With the notations of Prop \ref{proCRC}, $1_{\{ \omega'|\;\omega'(s)=\omega(s)\}}(\psi \circ i)^{\pi^r_s(\omega)}(\omega')=fX_{s_1}(\omega),...,X_{s_j}(\omega),X_{t_1}(\omega'))$. \begin{eqnarray} Q_{s,X_s(\omega)}((\psi \circ i)^{\pi^r_s(\omega)})=Q_{s,X_s(\omega)}(1_{\{\omega'|\; \omega'(s)=\omega(s)\}}(\psi \circ i)^{\pi^r_s(\omega)})\nonumber\\ =Q_{s,X_s(\omega)}(1_{\{\omega'|\; \omega'(s)=\omega(s)\}}(\phi_{\omega}(X_{t_1})) \label{eqscont} \end{eqnarray} where $\phi_{\omega}$ is the continuous function $\phi_{\omega}(z)=f(X_{s_1}(\omega),...,X_{s_j}(\omega),z)$. It follows from the Feller property that for given $x_1,...,x_j$, the map \begin{equation} y \rightarrow Q_{s,y}(f(x_1,...x_j,X_{t_1})\;\text{ is continuous} \label{eqyconti} \end{equation} We prove now that the map $\hat f : (x_1,...x_j,y) \rightarrow Q_{s,y}(f(x_1,...x_j,X_{t_1}))$ is continuous.\\The family of probability measures $(Q_{s,y})_{y \in \R^n}$ on $\Omega^s_T$ is weakly continuous thus for given $K>0$, $(Q_{s,y})_{||y|| \leq K}$ is tight. Thus for all $\epsilon>0$, there is a compact subset ${\cal K}$ of $\Omega^s_T$ such that for $||y|| \leq K$, $Q_{s,y}({\cal K}^c) < \epsilon$. $\cal K$ is compact for the Skorohod topology, thus from Theorem 12.3 of \cite{Bil} there is $A>0$ such that $\sup_{\omega' \in {\cal K}}|X_{t_1}(\omega')| \leq A$. Let $K'>0$. Due to the uniform continuity of $f$ on the compact set $\{||x_i||\leq K'\;\forall 1 \leq i \leq j,\; ||x_{t_1}|| \leq A\}$, being $Q_{s,y}$ a probability measure, for all $\epsilon>0$ there is $\eta>0$ such that for $||x_i|| \leq K'$, $||x'_i|| \leq K'$ and $||x_i-x'_i||<\eta$, \begin{equation}|Q_{s,y}(f(x_1,..,x_j,X_{t_1}))-Q_{s,y}(f(x'_1,..,x'_j,X_{t_1}))|<2(1+||f||) \epsilon, \;\;\forall y, \; \;||y|| \leq K \label{equnifc} \end{equation} The continuity of $\hat f:\;(x_1,...x_k,y) \rightarrow Q_{s,y}(g(x_1,...x_j,X_{t_1}))$ follows then easily from (\ref{eqyconti}) and (\ref {equnifc}). \item It follows from equation (\ref{eqscont}) and Proposition \ref{proCRC} that the map $\omega \rightarrow \hat f((X_{s_1}(\omega),...,X_{s_j}(\omega),X_s(\omega))$ is the canonical version of the $Q_{r,x}$ conditional expectation of $f_{(s_1,..,s_j),(t_1,...t_k)}$ given ${\cal B}^r_{s}$. \item By iteration. Assume that the result is satisfied for $k-1 \geq 1$. We prove it for $k$. It follows from the first step applied with $s=t_{k-1}$ that there is a continuous function $\psi$ on $\R^{j+(k-1)}$ such that $\psi_{(s_1,..,s_j),(t_1,...t_{k-1})}(\omega)$ is the canonical version of the $Q_{r,x}$ conditional expectation of $f_{(s_1,..,s_j),(t_1,...t_k)}$ given ${\cal B}^r_{t_{k-1}}$, which means that \\$T^{\theta}_{t_{k-1}}(f_{(s_1,..,s_j),(t_1,...t_k)}=\psi_{(s_1,..,s_j),(t_1,...t_{k-1})}$. The result follows then by iteration making use of the composition rule (\ref{eqcomp}). \end{itemize} \hfill $\square $ \begin{notation} Given $r\leq s$, denote ${\cal W}^r_s$ the algebra of continuous coordinates maps ${\cal B}^r_s$ measurable. \begin{equation} {\cal W}^r_s=\{f_{s_1,..,s_j}|r \leq s_1 \leq ...\leq s_j \leq s,\; j \in \N^*, f \in {\cal C}_b(\R^k)\} \label{eqW} \end{equation} where \begin{equation} f_{s_1,..,s_j}(\omega)=f((X_{s_1}(\omega),...,X_{s_j}(\omega)) \label{eqW2} \end{equation} \label{notcor2} \end{notation} From Proposition \ref{propDCC2}, we deduce the following Corollary: \begin{corollary} Assume that $\Theta$ satisfied hypothesis $H_{\theta}$ and has the Feller property. For all $Y \in {\cal W}^r_t$, $T^{\theta}_s(Y)$ belongs to ${\cal W}^r_s$. \label{cor2} \end{corollary} \section{Time Consistent Convex Continuous Dynamic Procedures} \label{secTCC} In this Section the Polish space is either ${\cal C}^r$ or ${\cal D}^r$. We assume that $\Theta$ satisfies hypothesis $H_{\theta}$ and has the Feller property. The goal of this Section is to construct time consistent convex procedures having a Feller property, associated to a stable set of probability measures ${\cal Q}^{\Theta}(\Gamma)_{r,y}$ generated by $\Theta (\Gamma)$ (cf Corollary \ref{corsta}). \begin{proposition} Let $F=\{f_{\alpha},\; \alpha \in A\}$ be a set of l.s.c. functions uniformly bounded from below on $(\R^n)^k$. \begin{itemize} \item There is a countable subset $\{f_{\alpha_n},\; n \in \N\}$ of $F$ such that \begin{equation} \sup_{n \in \N}f_{\alpha_n}=\sup_{\alpha \in A}f_{\alpha},\; \label{eqqs} \end{equation} \item If the set $F$ is a lattice upward directed, there is an increasing sequence $g_n$ of elements of $F$ such that \begin{equation} \sup_{\alpha \in A}f_{\alpha}=\lim_{n \rightarrow \infty}g_n,\; \label{eqqs2} \end{equation} \end{itemize} \label{lemmaqs} \end{proposition} {\bf Proof} \begin{enumerate} \item Let $(K_p)_{p \in \N}$ be the closed ball of radius $p$ in $(\R^n)^k$ centered in $0$. Let $(\phi_p)_{p \in \N^*}$ be an increasing sequence of continuous functions with compact support contained in $K_p$ such that the restriction of $\phi_p$ to $K_{p-1}$ is equal to $1$. \item Every function l.s.c. bounded from below is the limit of an increasing sequence of continuous bounded functions uniformly bounded from below. It follows then from {\it 1} that for all $\alpha$, there is a sequence $g^{\alpha}_p$ of continuous functions with compact support contained in $K_p$ such that $f_{\alpha}$ is the increasing limit of $g^{\alpha}_p$. Thus \begin{equation} \sup_{\alpha \in A} f_{\alpha}=\sup_{\alpha \in A,\; p \in \N^*} g^{\alpha}_p \label{eqsci} \end{equation} The set of continuous functions with compact support contained in $K_p$ is metrizable separable. It follows that the set $\{g^{\alpha}_p,\;\alpha \in A,\; p \in \N^*\}$ admits a countable dense subset. Thus there is a sequence $\alpha_n$ such that \begin{equation} \sup_{\alpha \in A,\; p \in \N^*} g^{\alpha}_p=\sup_{n \in \N,\; p \in \N^*} g^{\alpha_n}_p \label{eqsci2} \end{equation} By construction, for all $n,p$, $g^{\alpha_n}_p \leq f_{\alpha_n}$. Thus equation (\ref{eqqs}) follows from (\ref{eqsci}) and (\ref{eqsci2}). \item Let $f_{\alpha_n}$ in $F$ be such that equation (\ref{eqqs}) is satisfied. Define the sequence $g_n$ by iteration: $g_0=f_{\alpha_0}$ and choose for $g_{n+1}$ a function in $F$ such that $g_{n+1}\geq \sup(g_n,f_{\alpha_{n+1}})$. The sequence $g_n$ is by construction increasing. Let $g$ its limit, $\sup_{\alpha \in A}f_{\alpha}\geq g \geq \sup_{n \in \N} f_{\alpha_n}$. The result follows from (\ref{eqqs}). \end{enumerate} \hfill $\square $\\ We want now to construct general time consistent continuous procedures on ${\cal C}^r$ or ${\cal D}^r$. For this we introduce the notion of Feller property for a penalty. Recall that a penalty has been defined in Definition \ref{defpen}. \begin{definition} Let ${\cal Q}={\cal Q}^{\Theta}(\Gamma)_{r,y}$ be a stable set of probability measures on $(\Omega^r,{\cal B}^r_T)$ generated by $\Theta (\Gamma)$ (cf Corollary \ref{corsta}). A penalty $(\alpha_{s,t})_{r \leq s \leq t\leq T}$ defined on ${\cal Q}$ is a Feller penalty if \begin{equation} \forall \theta \in \Theta(\Gamma),\; \exists \beta^{\theta}_{st} \in {\cal C}(\R^n),\; \alpha_{s,t}(Q^{\theta}_{r,y})= \beta^{\theta}_{st}(X_s) \label{eqpenfe} \end{equation} \label{defpf} \end{definition} \begin{remark} The above Feller property is required only for the probability measures $Q^{\theta}_{r,y}$ generating the stable set ${\cal Q}$ but not for all the probability measures $Q^{\gamma}_{r,y}$ ($\gamma \in \Theta(\Gamma)^r_T$) in ${\cal Q}$. From the definition of a penalty (Definition \ref{defpen}), it follows that the function $\beta^{\theta}_{st}$ is bounded from above. We do not ask for boundedness. \label{remFell} \end{remark} \begin{notation} Let $0 \leq r \leq t$. Recall that ${\cal W}^r_t$ (Notation \ref{notcor2}) is the algebra of continuous coordinates functions. Denote also ${\cal L}_t$ the algebra ${\cal L}_t=\{f(X_t), \; f \text{continuous bounded on } \;\R^n \}$. In the following \begin{enumerate} \item ${\cal H}^r_t$ denotes either ${\cal W}^r_t$ or ${\cal L}_t$. \item $\hat {\cal H}^r_t$ denotes either $\hat {\cal W}^r_t$ or $\hat {\cal L}_t$, with \begin{equation} \hat {\cal W}^r_t=\{f(X_{t_1},...X_{t_k}),\; r \leq t_1 <..<t_k\leq t,\;\; \text{f lsc bounded from below}\} \label{eqwhat} \end{equation} \begin{equation} \hat {\cal L}_t=\{f(X_t)\; \text{f lsc bounded from below}\} \label{eqhhat} \end{equation} \end{enumerate} \label{notH} \end{notation} The following proposition is an adaptation to this general martingale setting of the ideas used in the proof of Theorem 4.8 of \cite{JBN-PDE}. \begin{proposition} Let $r \leq s \leq t$. Assume that $\Theta$ satisfies hypothesis $H_{\theta}$ and the Feller property. Let $\Gamma$ be a multivalued Borel mapping. Let $\alpha_{st}$ be a Feller penalty on the stable set of probability measures on $(\Omega^r,{\cal B}^r_T)$ generated by $\Theta (\Gamma)$. Let $Y \in \hat {\cal H}^r_t$, \begin{enumerate} \item For all $\theta \in \Theta(\Gamma)$, $T^{\theta}_s(Y)- \alpha_{s,t}(Q^{\theta}_{r,y})$ belongs to $ \hat {\cal H}^r_s$ \item For all $\gamma \in \Theta(\Gamma)^r_t$ and $s \in [r,t[$, there is $\eta_s \in \Theta(\Gamma)^r_t$ such that \begin{equation} T^{\gamma}_s(Y)- \alpha_{s,t}(Q^{\gamma}_{r,y})\leq T^{\eta_s }_s(Y)- \alpha_{s,t}(P^{\eta_s }_{r,y})=Z_s \label{eqcontin} \end{equation} and such that $T^{\eta_s }_s(Y)- \alpha_{s,t}(P^{\eta_s }_{r,y})=Z_s$ belongs to $ \hat {\cal H}^r_s$. \end{enumerate} \label{propc} \end{proposition} {\bf Proof} Notice first that for all $\gamma \in \Theta(\Gamma)^r_t$ the map $T^{\gamma}_s$ is defined on all functions ${\cal B}^r_t$ measurable bounded from below and that if $Y,Z$ are ${\cal B}^r_t$ measurable bounded from below, $T^{\gamma}_s(Y+Z)=T^{\gamma}_s(Y)+T^{\gamma}_s(Z)$. This follows from the linearity of $T^{\gamma}_s$ on bounded functions and from the continuity from below. \begin{enumerate} \item Let $\theta \in \Theta(\Gamma)$. Let $Y$ in ${\cal H}^r_t$, it follows from Proposition \ref{propDCC2} that $T^{\theta}_s(Y)$ belongs to ${\cal H}^r_s$ . If $Y$ belongs to $\hat {\cal H}^r_t$, $Y$ is the increasing limit of a sequence $Y_n \in {\cal H}^r_t$. From the continuity from below of $T^{\theta}_s$ (Corollary \ref{corext}), it follows that $T^{\theta}_s(Y)$ belongs to $ \hat {\cal H}^r_s$. It follows from the definition of a Feller penalty (Definitions \ref{defpf} and \ref{defpen}) that $-\alpha_{st}(Q^{\theta}_{r,y})$ belongs to $\hat {\cal H}^r_s$. This gives the result. \item $r=s_0<s_1<...<s_n=t$ be a subdivision associated to $\gamma$. For all $u \in [s_{n-1},s_n[$, $\gamma_{u}=\sum_{j \in I_{n-1}} 1_{A_{{n-1},j}}(\omega)\theta_{{n-1},j}(u,X_u(\omega))$. For all $j \in I_{n-1}$, and $s \in [s_{n-1},s_n[$, from 1., there is an element $Y_{s,j}$ of $ \hat {\cal H}^r_s$ such that $T^{\theta_{{n-1},j}}_{s}(Y)- \alpha_{s,t}(Q^{\theta_{{n-1},j}}_{r,y})=Y_{s,j}$. $Z_s=\sup_{j \in I_{n-1}}Y_{s,j}$ belongs to $ \hat {\cal H}^r_s$. Let $B_{s,j}$ the ${\cal B}^r_s$ measurable sets such that $Z_s=\sum_{j \in I_{n-1}}1_{B_{s,j}}Y_{s,j}$. Let $\eta_s \in \Theta(\Gamma)^r_t$ such that $\eta_s(u)(\omega)=\sum_{j \in I_{n-1}}1_{B_{s,j}}(\omega)\theta_{{n-1},j}(u,X_u(\omega))$ for all $u \in [s,t]$. $\eta_s$ satisfies the required conditions.\\ \item For $s \in [s_i,s_{i+1}[$ with $i+1<n$, the result is proved downward by induction. $\gamma_{s}=\sum_{j \in I_{i}} 1_{A_{i,j}}(\omega)\theta_{i,j}(u,X_u(\omega))$ Assume that $Z_v$ and $\eta_v$ satifying equation(\ref{eqcontin}) and $Z_v \in \hat {\cal H}^r_s$ have been constructed for all $v \geq s_{i+1}$. For all $j \in I_{i}$, $T^{\theta_{i,j}}_{s}(Z_{s_{i+1}})- \alpha_{s,t}(Q^{\theta_{i,j}}_{r,y})=\phi_{s,j}$ belongs to $ \hat {\cal H}^r_s$. Let $B_{s,j} \in {\cal B}^r_s$ such that $Z_s=\sup_{j \in I_{i}}\phi_{s,j}=\sum_{j \in I_{i}}1_{B_{s,j}}\phi_{s,j}$. Let $\eta_s \in \Theta(\Gamma)^r_t$ such that $\eta_s(u)=\eta_{s_{i+1}}(u)$ for all $u \geq s_{i+1}$ and $\eta_s(u)(\omega)=\sum_{j \in I_{i}}1_{B_{s,j}}\theta_{{i},j}(u,X_u(\omega))$ for all $u \in [s,s_{i+1}[$. $\eta_s$ and $Z_s$ satisfy the required conditions. \end{enumerate} \begin{theorem} Let $r \leq s \leq t \leq T$. Let $y \in \R^n$. Assume that $\Theta$ satisfies hypothesis $H_{\theta}$ and the Feller property. Let $\Gamma$ be a multivalued Borel mapping. Let ${\cal Q}={\cal Q}^{\theta}(\Gamma)_{r,y}$ be the stable set of probability measures generated by $\Theta(\Gamma)$. Let $(\alpha_{st})$ be a Feller penalty on ${\cal Q}$. Denote $T^{\gamma}_s=T^{Q^{\gamma}_{r,y}}_s$. For all $Y$ in $ \hat {\cal H}^r_t$, let \begin{equation} \Pi^{r,y}_{s,t}(Y)=\sup_{Q^{\gamma}_{r,y} \in {\cal Q}^{\theta}(\Gamma)_{r,y}} (T^{\gamma}_s(Y)- \alpha_{s,t}(Q^{\gamma}_{r,y})) \label{eqtcd0} \end{equation} \begin{enumerate} \item For all $Y \in \hat {\cal H}^r_t$. $\Pi^{r,y}_{s,t}(Y)$ belongs to $ \hat {\cal H}^r_s$. Moreover for all $s \in [r,t]$, there is a sequence $Q_n=Q^{\gamma_n}_{r,y}$ in ${\cal Q}$, $\gamma_n$ in $\Theta(\Gamma)^r_t$ such that $T^{\gamma_n}_s(Y)-\alpha_{s,t}(Q^{\gamma_n}_{r,y})$ belongs to $\hat {\cal H}^r_s$, and such that $\Pi^{r,y}_{s,t}(Y)$ is the increasing limit: \begin{equation} \Pi^{r,y}_{s,t}(Y)=\lim [T^{\gamma_n}_s(Y)- \alpha_{s,t}(Q^{\gamma_n}_{r,y})] \label{eqtcd1} \end{equation} \item $\Pi^{r,y}_{s,t}$ is a convex monotone map continuous from below on $ \hat {\cal H}^r_t$ with values in $ \hat {\cal H}^r_s$.\\ \item For all $r \leq s \leq t \leq u$, for all $Y \in \hat {\cal H}^r_u$, \begin{equation} \Pi^{r,y}_{s,u}(Y)=\Pi^{r,y}_{s,t}(\Pi^{r,y}_{t,u}(Y)) \label{tc510} \end{equation} \item In case ${\cal H}^r_t={\cal L}_t$, the sequence $\gamma_n$ of \it{1.} can be choosen independently of $r$. \end{enumerate} \label{ThmTCCP} \end{theorem} {\bf Proof} \begin{enumerate} \item Let $Y \in \hat {\cal H}^r_t$. For $Q^{\gamma}_{r,y}\in {\cal Q}$, let $Y^{\gamma}_s=T^{\gamma}_s(Y)- \alpha_{s,t}(Q^{\gamma}_{r,y})$. Let ${\cal Z}=\{Y^{\gamma}_s\; Q^{\gamma}_{r,y}\in {\cal Q},\; \gamma \in \Theta(\Gamma)^r_t\; |\;Y^{\gamma}_s \in \hat {\cal H}^r_s\}$. It follows from the stability by bifurcation of ${\cal Q}$, from the expression of $T^{\gamma}$ (equation (\ref{Tgamma})) and from the locality of the penalty that ${\cal Z}$ is a lattice upward directed. \begin{itemize} \item If ${\cal H}^r_t={\cal W}^r_t$, there is a lsc function $f$ bounded from below such that $Y=f(X_{s_1},X_{s_2},..X_{s_l})$. Let $j$ such that $s_j < s \leq s_{j+1}$. From Proposition \ref {propDCC2} for all $\theta \in \Theta(\Gamma)$, $Y^{\theta}_s=g(X_{s_1},..., X_{s_j},X_s)$ for some function $g$ continuous bounded from below on $(\R^n)^{j+1}$. It follows from the proof of Proposition \ref{propc} that $\{g\;|g(X_{s_1},..., X_{s_j},X_s) \in {\cal Z}\}$ is a lattice of lsc functions bounded from below on $(\R^n)^{j+1}$. The result follows from Proposition \ref{lemmaqs}. \item In case ${\cal H}^r_t={\cal L}_t$, $Y=f(X_t)$. For all $\gamma \in \Theta(\Gamma)^r_T$, $Y^{\gamma}_s=g(X_s)$ for some function $g$ Borelian bounded from below on $(\R^n)^{j+1}$ independent of $r$. A similar proof as the above one proves {\it 1.} and {\it 4.} \end{itemize} \item From the linearity of $T^{\gamma}_s$ on bounded functions and the continuity form below, it follows that for all non negative real numbers $c$ and $d$, , for all $Y,Z$ in ${\cal H}^r_t$, $T^{\gamma}_s (cY+dZ)= c T^{\gamma}_s (Y)+dT^{\gamma}_s (Z)$. The convexity of $\Pi^{r,y}_{s,t}$ follows. Let $Y_n$ be increasing to $Y$. Every $T^{\gamma}_s$ is continuous from below thus $\Pi^{r,y}_{s,t}( Y)= \sup_{\gamma}T^{\gamma}_s(\lim Y_n)=\sup_{\gamma,n}T^{\gamma}_s(Y_n)=\sup_n\Pi^{r,y}_{s,t}(Y_n)$. \item Let $Y \in \hat {\cal H}^r_u$. From 1., $\Pi^{r,y}_{t,u}(Y)$ belongs to $\hat {\cal H}^r_t$ and we have the following increasing limits: \begin{equation} \Pi^{r,y}_{s,t}(\Pi^{r,y}_{t,u}(Y))=\lim_{n \rightarrow \infty} [T^{\gamma_n}_s(\Pi^{r,y}_{t,u}(Y))- \alpha_{s,t}(Q^{\gamma_n}_{r,y})] \label{eqtcd2} \end{equation} \begin{equation} \Pi^{r,y}_{t,u}(Y)=\lim_{k \rightarrow \infty} [T^{\delta_k}_t((Y))- \alpha_{t,u}(Q^{\delta_k}_{r,y})] \label{eqtcd3} \end{equation} Fom the continuity from below of $T^{\gamma_n}_s$, it follows that \begin{equation} \Pi^{r,y}_{s,t}(\Pi^{r,y}_{t,u}(Y))=\sup_{n,k} (T^{\gamma_n}_s[T^{\delta_k}_t((Y))- \alpha_{t,u}(Q^{\delta_k}_{r,y})]- \alpha_{s,t}(Q^{\gamma_n}_{r,y})) \label{eqtcd4} \end{equation} From the stability of ${\cal Q}$, for all $n$ and $k$, let $\lambda_{n,k}$ be defined as in Theorem \ref{thmsta} i), such that $T^{\gamma_n}_s(\phi)=T^{\lambda_{n,k}}_s(\phi)$ for all $\phi$ ${\cal B}^r_t$ measurable bounded and $T^{\delta_k}_t=T^{\lambda_{n,k}}_t$. From the chain rule, $T^{\gamma_n}_s \circ [T^{\delta_k}_t]= T^{\lambda_{n,k}}_s$. From the definition of $\lambda_{n,k}$, the local property of the penalty and the cocycle condition, it follows that $T^{\gamma_n}_s[- \alpha_{t,u}(Q^{\delta_k}_{r,y})]- \alpha_{s,t}(Q^{\gamma_n}_{r,y})=- \alpha_{s,u}(Q^{\lambda_{n,k}}_{r,y})$. Thus from (\ref{eqtcd4}), it follows that \begin{equation} \Pi^{r,y}_{s,t}(\Pi^{r,y}_{t,u}(Y))=\sup_{n,k} [ T^{\lambda_{n,k}}_s(Y))- \alpha_{s,u}(Q^{\lambda_{n,k}}_{r,y})] \leq \Pi^{r,y}_{s,u}(Y) \label{eqtcd5} \end{equation} Conversely, $\Pi^{r,y}_{s,u}(Y)$ is the increasing limit of $T^{\nu_j}_s(Y)- \alpha_{s,u}(Q^{\nu_j}_{r,y})$. Making use of the chain rule for $T^{\nu_j}$ and of the cocycle condition, we get the inequality $\Pi^{r,y}_{s,u}(Y) \leq \Pi^{r,y}_{s,t}(\Pi^{r,y}_{t,u}(Y))$. \item If $Y \in {\cal L}_t$, $Y=f(X_t)$. \end{enumerate} \begin{definition} Let $r \leq s \leq t \leq T$. Let $\Theta$ satisfying hypothesis $H_{\theta}$. let ${\cal Q}={\cal Q}^{\theta}(\Gamma)_{r,y}$ be the stable set of probability measures generated by $\Theta(\Gamma)$. Let $(\alpha_{st})$ be a Feller penalty on ${\cal Q}$. For all $Y$ in $ \hat {\cal H}^r_t$, let \begin{equation} \Pi^{r,y}_{s,t}(Y)=\sup_{Q^{\gamma}_{r,y} \in {\cal Q}^{\theta}(\Gamma)_{r,y}} (T^{Q^{\gamma}_{r,y}}_s(Y)- \alpha_{s,t}(Q^{\gamma}_{r,y}) \label{eqtcd6} \end{equation} One says that $\Pi^{r,y}_{s,t}$ is a time consistent dynamic procedure on $ \hat {\cal H}^r_t$ if all properties of Theorem \ref{ThmTCCP} are satisfied. \label{defTCPP} \end{definition} \begin{proposition} Let ${\cal P}={\cal Q}^{\theta}(\Gamma)_{r,y}$. Denote $\overline {\cal W}^r_t$ the closure of the lattice vector space ${\cal W}^r_t$ for the norm $\sup_{Q \in {\cal P}}E_Q(|X|)$. Assume that for all $s$, and $X \in {\cal W}^r_t$ , $\Pi^{r,y}_{s,t}(X)$ belongs to $\overline {\cal W}^r_s$. The procedure $\Pi^{r,y}_{s,t}$can be uniquely extended to $\overline {\cal W}^r_t$. The time consistency (equation \ref{tc510}) extends to $\overline {\cal W}^r_t$.\\ In case where ${\cal P}$ is a weakly relatively compact set of probability measures on ${\cal C}([r,T],\R^n)$, $\overline {\cal W}^r_t$ is equal to $L^1(c)$. \end{proposition} {\bf Proof} Let $X,Y$ in ${\cal W}^r_t$. From equation (\ref{eqtcd6}), $\Pi^{r,y}_{s,t}(Y)$ is the increasing limit of $T^{\gamma_n}_s(Y)- \alpha_{s,t}(Q^{\gamma_n}_{r,y})$. It follows that $\Pi^{r,y}_{st}(Y) \leq \Pi^{r,y}_{st}(X)+ sup_{n \in \N}T^{\gamma_n}_s(Y-X)$. Changing the roles of $X$ and $Y$, it follows that \begin{equation} |\Pi^{r,y}_{st}(Y)-\Pi^{r,y}_{st}(X)| \leq \sup_{n,k}[T^{\gamma_n}_s(|Y-X|), T^{\delta_k}_s(|Y-X|)] \label{eqextens} \end{equation} As in the proof of equation (\ref{eqtcd1}), it follows from Proposition \ref{lemmaqs} that \\$\sup_{Q^{\gamma} \in {\cal P}}T^{\gamma}_s(|X-Y|)$ is the increasing limit of a sequence $T^{\alpha_n}_s (|X-Y|)$. It follows from equation (\ref {eqextens}) and the stability of ${\cal P}$ that \begin{equation} \sup_{P \in {\cal P}} E_P(|\Pi^{r,y}_{st}(Y)-\Pi^{r,y}_{st}(X)|) \leq \sup_{P \in {\cal P}} E_P(|X-Y|) \label{eqL1C} \end{equation} It follows from equation (\ref{eqL1C}), that $\Pi^{r,y}_{st}$ can be uniquely extended to $\overline {\cal W}^r_t$, with values in $\overline {\cal W}^r_s$, equation (\ref{tc510}) is then satisfied on $\overline {\cal W}^r_t$.\\ The last result follows from Lemma \ref{lemmacont}. \section{ Construction of Feller penalties} \label{secTCCC} \subsection{Penalties for continuous diffusions} \label{secTCCC0} In this Section we restrict to the case of continuous paths. We consider the martingale problem for diffusions. In \cite{JBN-PDE} we have constructed Feller penalties associated to probability measures solution to the martingale problem for continuous diffusions. In \cite{JBN-PDE} the probability measures in the stable set were all solutions to a martingale problem with the same function $a(t,x)$. In the present paper the probability measures are no more equivalent, they are even not dominated. The construction of the penalties that we describe below is the generalization of the construction that we have done in \cite{JBN-PDE}. The ideas are the same. In this paper we chose the parameters $a$ and $b$. \begin{definition} $\Theta$ is the set of continuous bounded functions $\theta(t,x)=$\\ $(a(t,x),b(t,x)))$ with values in $M_n(\R)) \times \R^n$ such that for all $(t,x)$ the matrix $a(t,x)$ is invertible. Let $\Gamma$ be a multivalued Borel mapping from $\R_+ \times \R^n$ to $M_n(\R) \times \R^n$, $\Theta(\Gamma)=\{\theta \in {\Theta},\; \Gamma\;\text{valued}\}$. \label{deftetacont} \end{definition} For general $\gamma=(\eta,\mu) \in {\Theta}^r_T$ we denote $P^{\eta,\mu}_{r,y}$ the unique solution to the martingale problem starting from $y$ at time $r$ constructed previously. In the case $\gamma \in \Theta$, $\gamma(u,\omega)=\theta(u,X_u(\omega))$ for some continuous $\theta(u,x)$. When we want to emphasize the fact that we refer to continuous parameters $\eta$ and $\mu$, the probability measure $P^{\eta,\mu}_{r,y}$ will be denoted $Q^{a,b}_{r,y}$. As in \cite{JBN-PDE} we introduce hypothesis $H_g$. \begin{definition} Hypothesis $H_g$\\ \begin{enumerate} \item $g:\R_+ \times \R^n \times (M_n(\R) \times \R^n) \rightarrow \R\cup {\infty}$ is a ``Caratheodory function on $\Gamma$''\\ More precisely, $g$ is Borelian and for all $u$, the restrition of $g_u$ to $\{(x,y), y \in \Gamma(u,x)\}$ is continuous ($g_u(x,y)=g(u,x,y)$). \item $g$ has polynomial growth on $\Gamma$ \begin{equation} \forall (t,x,y), \; y \in \Gamma(t,x),\;\;|g(t,x,y)| \leq C(1+||x||^m) \label{eqpolg} \end{equation} \item $g$ is bounded from above on $\Gamma$ which means that $g$ is bounded from above on $\{(t,x,y), \; y \in \Gamma(t,x)\}$ \end{enumerate} \label{defHgc} \end{definition} This last condition on $\Gamma$ which was not present in \cite{JBN-PDE} could be suppressed with a little more work.\\ Recall that from Corollary \ref{corsta}, the set ${\cal Q}^{\Theta}(\Gamma)_{r,y}=\{Q^{\gamma}_{r,y},\; \gamma \in {\Theta}(\Gamma)^r_T\}$ is stable. \begin{proposition} Let $\Gamma$ be a multivalued Borel mapping. Let $\Theta(\Gamma)$ be as in Definition \ref{deftetacont}. Assume that $g$ satisfies Hypothesis $H_g$. Let $t>0$. \begin{enumerate} \item For all $\theta=(a,b) \in \Theta(\Gamma)$, there is a real valued continuous map $(s,x) \in [0,t]\times \R^n \rightarrow L^{a,b}_t(g)(s,x)$ such that \begin{equation} Q^{a,b}_{s,y}(\int_s^t g(u, X_u,\theta (u,X_u)) du )=L^{a,b}_t(g)(s,y) \;\;\;\forall s \in [0,t]\;\;and\; y \in \R^n \label{eqFell47-3} \end{equation} \item For all $0 \leq r \leq s \leq t$ and all $y \in \R^n$, for $Q=Q^{a,b}_{r,y}$, \begin{equation} T^{Q}_s[-\int_s^t g(u, X_u,\theta(u,X_u)) du ]=-L^{a,b}_t(g)(s,X_s(\omega)) \label{eqFell47b-3} \end{equation} \end{enumerate} Notice that for all $y \in \R^n$ $L^{a,b}_t(g)(t,x)=0$ \label{propcont3} \end{proposition} {\bf Proof} {\it 1.} is an application of the first statement of Proposition 4.2 of \cite{JBN-PDE}\\ {\it 2.} $-g$ is bounded from below, so making use of the extension of $T^{Q}_s$ defined on Corollary \ref{corext} it follows that \begin{eqnarray} T^{Q}_s[-\int_s^t g(u, X_u,\theta(u,X_u)) du ]&=&Q^{a,b}_{s,X_s(\omega)}[-\int_s^t g(u, X_u,\theta(u,X_u)) du ]\nonumber\\ &=&- L^{a,b}_t(g)(s,X_s(\omega)) \label{eqfepe} \end{eqnarray} We define now the penalty of every probability measure $Q^{\gamma}_{r,y}$ in ${\cal Q}^{\Theta}(\Gamma)_{r,y}$. \begin{definition} For all $Q^{\gamma}_{r,y}$, for all $r \leq s\leq t \leq T$, we define the penalty \begin{equation} \alpha_{s,t}(Q^{\gamma}_{r,y})=-T^{\gamma}_s[-\int\limits_{s}^{t} g(u,X_u(\omega),\gamma(u,\omega)) d u] \label{eqP1_0} \end{equation} \label{defpen0} \end{definition} \begin{corollary} The above definition provides a Feller penalty according to Definitions \ref{defpen} and \ref{defpf}. \label{corpen} \end{corollary} {\bf Proof} It follows from hypothesis $H_g$ that for all $r \leq s \leq t$, the function $-\int\limits_{s}^{t} g(u,X_u(\omega),\gamma(u,\omega)) d u]$ is ${\cal B}^r_t$ measurable bounded from below, thus from Corollary \ref{corext}, $\alpha_{s,t}(Q^{\gamma}_{r,y})$ is ${\cal B}^r_s$ measurable bounded from above. The cocycle condition for $\alpha_{s,t}(Q^{\gamma}_{r,y})$ follows from the chain rule for $T^{\gamma}$ ( Corollary \ref{corext})and the additivity of $T^{\gamma}_s$ on functions bounded from below. The local condition follows from the Definition of the penalty. The Feller property of the penalty follows from Proposition \ref{propcont3}. \subsection{Penalties for diffusions with Levy generators} \label{secpenlevy} In this Section we consider probability measures on the Polish space ${\cal D}^r$ of c\`adl\`ag paths endowed with the Skorhokod topology. The notations are those of Section \ref{secDiLev}. \begin{definition} \begin{enumerate} \item Let $C>0$. ${\cal M}_C$ denotes the set of $\sigma$-finite measures $\mu$ on $\R^n -\{0\}$ such that \begin{equation} \int [||y||^2 1_{||y|| \leq 1}+||y||1_{||y|| >1}]\mu(dy) \leq C \label{eqMc} \end{equation} \item $\Theta$ is the set of $(a,b,M)$ such that \begin{itemize} \item $a,b$ are continuous bounded functions on $\R_+ \times \R^n$, with values in $M^n(\R)$, $\R^n$, and for all $(t,x)$ $a(t,x)$ is invertible. \item $M$ satisfies hypothesis $M_C$ (Definition \ref{defMc}), i.e. for all $(t,x)$, $M(t,x) \in {\cal M}_c$. Furthermore for all Borelian subset $\Delta$ of $\R^n-\{0\}$, \begin{equation} \int _{\Delta}\frac{||y||^2}{1+||y||^2} M(s,x,dy)\; \label{eqbound1} \end{equation} is a continuous bounded function of $(s,x)$. \end{itemize} \item ${\cal M}_c$ is endowed with the weak topology. Let $\Gamma$ be a closed convex multivalued Borel mapping from $\R_+ \times \R^n$ to $M_n(\R) \times \R^n \times {\cal M}_c$. Let $\Theta(\Gamma)=\{\theta \in {\Theta},\; \Gamma\;\text{valued}\}$. \end{enumerate} \label{deftetacont2} \end{definition} We want now to define the penalty of every probability measure $Q^{\gamma}_{r,y}$ in ${\cal Q}^{\Theta}(\Gamma)_{r,y}$. Recall that $\Theta^r_T$ is described by Definition \ref{Slambda}. For all $\gamma \in \Theta(\Gamma)^r_T$, write $\gamma(u,\omega)=(\eta,\lambda,\mu)(u,\omega)$, where $\eta$ takes values in $M_n(\R)$, $\lambda$ in $\R^n$ and $\mu$ takes values in ${\cal M}_C$. \begin{definition} Let $g$ be a real valued function on $\R^+ \times M_n(\R)\times \R^n \times {\cal M}_C$ such that for all $\theta=(a,b,M) \in \Theta(\Gamma)$, $ g^{\theta}(u,x)=g(u,x,a(u,x),b(u,x),M(u,x)$ is continuous bounded. We define the penalty \begin{equation} \alpha_{s,t}(Q^{\gamma}_{r,y})=T^{\gamma}_s[\int\limits_{s}^{t} g(u,X_u(\omega),\eta(u,\omega),\lambda(u,\omega),\mu(u,\omega))d u] \label{eqP1_} \end{equation} \label{defpenlev} \end{definition} \begin{proposition} The above definition provides a penalty according to Definition \ref{defpen}. The penalty is a Feller penalty (Definition \ref{defpf}). More precisely, for all $\theta=(a,b,M) \in \Theta(\Gamma)$, there is a continuous function $L_{t}^{\theta}(g)$ on $[0,t] \times \R^n$ such that $\forall s \in [0,t]\;\;and\; y \in \R^n$, \begin{equation} Q^{a,b,M}_{s,y}(\int_s^t g(u,X_u(\omega),a(u,X_u(\omega)),b(u,X_u(\omega)), M(u,X_u(\omega))du= L_{t}^{\theta}(g)(s,y) \;\;\; \label{eqFell47-4} \end{equation} \item For all $0 \leq r \leq s \leq t$ and all $y \in \R^n$, for $Q=Q^{a,b,M}_{r,y}$, \begin{equation} T^{Q}_s[\int_s^t g(u,X_u(\omega),a(u,X_u(\omega)),b(u,X_u(\omega)), M(u,X_u(\omega))du]=L_{t}^{\theta}(g)(s,X_s(\omega)) \label{eqFell47b-4} \end{equation} \label{proppen2} \end{proposition} {\bf Proof} \begin{enumerate} \item It follows easily from the hypothesis, and the description of $\Theta(\Gamma)^r_T$ (cf Definition \ref{Slambda}) that for all $\gamma=(\eta,\lambda,\mu) \in \Theta(\Gamma)^r_T$, the function $\int\limits_{s}^{t} g(u,X_u(\omega),\eta(u,\omega),\lambda(u,\omega),\mu(u,\omega))$ is ${\cal B}^r_t$- measurable bounded. From Proposition \ref{propcanreg} it follows that $\alpha_{s,t}(Q^{\gamma}_{r,y})$ is ${\cal B}^r_s$ measurable bounded. \item The cocycle condition follows easily from the definition of the penalty (\ref{eqP1_}), the linearity for $T^{\gamma}$ and the chain rule for $T^{\gamma}$. \item locality Let $\gamma,\eta \in {\Theta^r_T}$, for all $A \in {\cal B}^r_s$, if $1_A \gamma(u,\omega)=1_A \eta(u,\omega)$, for all $u \in]s,t]$. from Proposition \ref{propcanreg} equation (\ref{Tgamma}), $T^{\gamma}_s$ and $T^{\eta}_s$ coincide on bounded variables ${\cal B}^r_t$ measurable. The locality follows then from the definition of the penalty. \item Let $\theta \in \Theta(\Gamma)$, $\theta=(a,b,M)$, $a,b$ are continuous bounded functions of $(u,x)$, and $M$ satisfies hypothesis $M_C$. Let $\epsilon>0$. Let $K$ such that equation (\ref{eqLM0}) is satisfied for all $0 \leq r \leq t$ and $||y|| \leq D$. The function $g^{\theta}$ is bounded and uniformly continuous on $[0,T] \times \{||x|| \leq K\}$. Making use of equation (\ref{eqLM}) and of arguments similar to those of Proposition \ref{lemmFell}, it follows that for all $s' \leq t$, there are $s' \leq t_1<t_2<...<t_k \leq t$ and a continuous bounded function $f_{s',t}$ on $(\R^n)^k$ such that for all $||y|| \leq D$, for all $0 \leq r \leq s'$, \begin{equation} Q^{a,b,M}_{r,y}(|\int\limits_{s'}^{t} g^{\theta}(u,X_u(\omega))du-f_{s',t}(X_{t_1},X_{t_2},..X_{t_k})|<\epsilon \label{equnif00} \end{equation} From Proposition \ref{lemmFell}, there is a continuous function $\hat f_{s',t}$ on $[0,s'] \times \R^n$ such that $Q^{a,b,M}_{r,y}(f_{s',t}(X_{t_1},X_{t_2},..X_{t_k}))=\hat f_{s',t}(r,y)$. It follows from (\ref{equnif00}) that $Q^{a,b,M}_{r,y}(\int\limits_{s}^{t} g^{\theta}(u,X_u(\omega))du$ is the uniform limit on $[0,s'] \times \{||y|| \leq D\}$ of a sequence of continuous functions. It is thus continuous. The function $g$ being bounded, the continuity on $[0,t] \times \R^n$ of the function $L^{\theta}_t$ defined by equation (\ref{eqFell47-4}) follows then easily. Equation (\ref{eqFell47b-4}) follows from Proposition \ref{proCRC}. \end{enumerate} \section{Feller property of the time consistent dynamic procedure} In all the following we restrict to the case of continuous diffusions or diffusions with Levy generator. The notations and hypothesis are those of Section \ref{secTCCC}. We prove now that with the specific choice for the penalty which was made in Subsections \ref{secTCCC0} and \ref{secpenlevy}, the procedures $\Pi^{r,y}_{s,t}$ that we have constructed for all given $r$ and $y$ are strongly connected and have a Feller property. We introduce now a larger set of parameters: $\tilde \Theta(\Gamma)^r_T$. The definition of $\tilde \Theta(\Gamma)^r_T$ is similar to the definition of $\Theta(\Gamma)^r_T$ (cf Definition \ref{Slambda}), but now the set $I_0$ is a finite set and not necessarily a singleton. Every element $\tilde \gamma$ of $\tilde \Theta(\Gamma)^r_T$, satisfies equation (\ref{eqstable}). For $\tilde \gamma \in \tilde \Theta(\Gamma)^r_T$, the process $(Z^{\tilde \gamma}_{r,u})_{r \leq u \leq T}$ is defined by equation (\ref{eqstable2}). \begin{remark} Let $0 \leq s <t$. Let $x \in \R^n$. Let $\tilde \gamma \in \tilde \Theta(\Gamma)^s_t$. The sets $A_{0,j}, j \in I_0$ belong to the $\sigma$-algebra generated by $X_s$ and form a partition of $\Omega^r_t$. It follows that there is $j_x \in I_0$ such that $\{X_s(\omega)=x\} \subset A_{0,j_x}$. Therefore a probability measure $Q$ on $(\Omega^s,{\cal B}^s_t)$ is solution to the martingale problem $(Z^{\tilde \gamma}_{s,u})_{s \leq u \leq t}$ starting from $x$ at time $s$ (Definition \ref{defFellprop}) if and only if $Q$ is solution to the martingale problem $(Z^{\gamma_x}_{s,u})_{s \leq u \leq t}$ starting from $x$ at time $s$, where $(\gamma_x)(u,\omega)=\theta_{0,j_x}(u,X_u(\omega))$ for all $s \leq u < s_1$, and $(\gamma_x) (u,\omega)=\tilde\gamma(u,\omega)$ for all $s_1 \leq u$. Thus ${\gamma_x}$ belongs to $\Theta(\Gamma)^s_t$. \label{remtheta0} \end{remark} The following proposition is a more precise version of Proposition \ref{propc}. \begin{proposition} Let $h$ be a continuous function on $\R^n$ bounded from below. For all $\gamma \in \Theta(\Gamma)^r_t$ there is a continuous function $h^{\gamma}$ on $[0,t] \times \R^n$ bounded from below and for all $r \leq s \leq t$, there is $\eta_s \in \tilde \Theta(\Gamma)^s_t$ such that \begin{enumerate} \item \begin{equation} \forall x \in \R^n,\;T^{P^{\eta_s}_{s,x}}_s(h(X_t))-\alpha_{s,t}(P^{\eta_s}_{s,x})=h^{\gamma}(s,x) \label{eqetas} \end{equation} There is a partition $B_j$ of $\R^n$ in Borelian sets, and $\eta^j_s \in \Theta(\Gamma)^s_t$ such that \begin{equation} \forall x \in B_j,\;T^{P^{\eta^j_s}_{s,x}}_s(h(X_t))-\alpha_{s,t}(P^{\eta^j_s}_{s,x})=h^{\gamma}(s,x) \label{eqetasj} \end{equation} \begin{equation} \forall x \in B_j,\;h^{\eta^j_s}(u,x)=h^{\gamma}(u,x), \; \forall s \leq u \leq t \label{eqetasj2} \end{equation} \item For all $\eta \in \theta(\Gamma)^r_t$ whose restriction to $[s,t]$ is equal to $\eta_s$, \begin{eqnarray} \forall y \in \R^n,\;\;T^{P^{ \eta}_{r,y}}_s (h(X_t)-\alpha_{s,t}(P^{\eta}_{r,y})=h^{\gamma}(s,X_s)\nonumber\\ T^{Q^{\gamma}_{r,y}}_s(h(X_t))-\alpha_{s,t}(Q^{\gamma}_{r,y}) \leq T^{P^{\eta}_{r,y}}_s(h(X_t))-\alpha_{s,t}(P^{ \eta}_{r,y})\nonumber\\ h^{\gamma}(u,x)=h^{\eta}(u,x) \;\;\forall s \leq u \leq t \label{eqnfell} \end{eqnarray} \end{enumerate} \label{propsup} \end{proposition} {\bf Proof} \begin{itemize} \item Step 1: Let $\theta \in \Theta(\Gamma)$, let $h^{\theta}$ be the function defined on $[0,t] \times \R^n$ by \begin{equation} \;T^{P^{\theta}_{s,x}}_s(h(X_t))-\alpha_{s,t}(P^{\theta}_{s,x})=h^{\theta}(s,x) \label{eqhteta} \end{equation} The continuity of the map $h^{\theta}$ on $[0,t] \times \R^n$ follows from Theorem 7.1 of \cite{SV1} (see Proposition 2.7 of \cite{JBN-PDE} for details), and from Proposition \ref{propcont3} in case of diffusions. In case of diffusions with Levy generator, the continuity of $h^{\theta}$ follows from Propositions \ref{lemmFell} and \ref{proppen2}. Equation \begin{equation} \forall 0 \leq r \leq s \leq t, \forall y \in \R^n,\;\;T^{P^{\theta}_{r,y}}_s (h(X_t))-\alpha_{s,t}(P^{\theta}_{r,y})=h^{\theta}(s,X_s) \end{equation} follows then from Lemma \ref{lemmaCC2}, Propositions \ref{propcont3} and \ref{proppen2}. \\ \item Step 2: The proof of the existence of $h^{\gamma}$ and $\eta_s$ satisfying (\ref{eqnfell}) follows the proof of Theorem 4.8 in \cite{JBN-PDE}. We sketch the proof. Let $r=s_0<s_1<...<s_n=t$ be a subdivision associated to $\gamma$. For $u \in [s_i,s_{i+1}[$, $\gamma(u,\omega)=\sum_{j \in I_i} 1_{A_{i,j}}(\omega)\theta_{i,j}(u,X_u(\omega))$, $\theta_{i,j} \in {\Theta}(\Gamma)$. $h^{\gamma}$ and $\eta_s$ are defined recursively. Assume that $h^{\gamma}$ is defined and continuous on $[s_{i+1},t] \times \R^n$, and that $\eta_u$ is defined for $u \in [s_{i+1},t]$. Let $h_{i,j}$ be the continuous function on $[s_i,s_{i+1}]$ such that $\forall s \in [s_i,s_{i+1}]$, \begin{equation} \forall x \in \R^n,\;T^{P^{\theta_{i, j}}_{s,x}}_s(h^{\gamma}(s_{i+1},X_{s_{i+1}}))-\alpha_{s,s_{i+1}}(P^{\theta_{i,j}}_{s,x})=h_{i,j}(s,x) \label{eqhteta2} \end{equation} For $s_i \leq s < s_{i+1}$, let $h^{\gamma}(s,x) =\sup_{j \in I_i} h_{i,j}(s,x)$. Given $s \in [s_i, s_{i+1}[$, there is a Borelian partition $B_{j},j \in I_i$ of $\R^n$ such that $h^{\gamma}(s,x) =\sum _{j \in I_i} 1_{B_{j}}(x) h_{i,j}(s,x)$. Let $\eta_s(u, \omega)=\eta^j_s(u, \omega)=\eta_{s_{i+1}}(u, \omega)$ for $u \geq s_{i+1}$. Let $\eta^j_s(u,\omega)=\theta_{i,j}(u,X_u(\omega))$ and $\eta_s(u,\omega)= \sum_j1_{B_{j}}(X_s)\theta_{i,j}(u,X_u(\omega))$ for $s \leq u <s_{i+1}$. We end the proof of {\it 1.} making use of the expression of $T^{Q^{\gamma}_{r,y}}_s$, cf Theorem \ref{propcanreg} and of Remark \ref{remtheta0}. Notice that $\eta_r$ belongs to $\Theta(\Gamma)^r_t$, $\eta^j_s \in \Theta(\Gamma)^s_t$ and $\eta_s \in \tilde \Theta(\Gamma)^s_t$ \item Step 3: It follows from Step 1, from the expressions of $T^{Q^{\gamma}_{r,y}}_s$, $T^{P^{\eta}_{r,y}}_s$, and the properties of the penalty that the construction given above leads to $h^{\gamma}$ and $\eta_s$ such that equation (\ref{eqnfell}) is satisfied for all $\eta \in \theta(\Gamma)^r_t$ whose restriction to $[s,t]$ is equal to $\eta_s$. \end{itemize}\hfill $\square $ \begin{theorem} \begin{enumerate} \item Let $0 \leq r \leq T$. Let $\hat {\cal W}^r_T$ be the cone of coordinate functions $f(X_{t_1},..X_{t_k})$, $f$ lower semi continuous bounded from below (Notation \ref{notH}). Let $(\Pi^{r,y}_{s,t})_{r \leq s \leq t \leq T}$ be defined on $\hat {\cal W}^r_T$ by \begin{equation} \Pi^{r,y}_{s,t}(Y)=\sup_{Q^{\gamma}_{r,y} \in {\cal Q}^{\theta}(\Gamma)_{r,y}} (T^{Q^{\gamma}_{r,y}}_s(Y)- \alpha_{s,t}(Q^{\gamma}_{r,y})) \label{eqtcd7} \end{equation} where $\alpha_{s,t}$ is the penalty constructed in Section \ref{secTCCC}. Then $(\Pi^{r,y}_{s,t})_{r \leq s \leq t \leq T}$ is a time consistent convex dynamic procedure on $\hat {\cal W}^r_t$. \item It has the following Feller property: For all $h$ lower semi continuous function on $\R^n$ bounded from below, there is a lower semi-continuous function $\tilde h$ bounded from below on $[0,t] \times \R^n$ such that \begin{equation} \forall x \in \R^n,\;\Pi^{s,x}_{s,t}(h(X_t))=\tilde h (s,x) \label{eqmeas20} \end{equation} \begin{equation} \forall 0 \leq r < s \leq t, \;\forall y \in \R^n,\;\;\Pi^{r,y}_{s,t}(h(X_t))=\tilde h (s,X_{s}) \label{eqmeas0} \end{equation} Furthermore for all given $r$ and $y$, for all $r<s \leq t$, there is a sequence $\eta_n \in \Theta(\Gamma)^r_t$ such that $\Pi^{r,y}_{s,t}(h(X_t))$ is the increasing limit of $(T^{P^{\eta_n}_{r,y}}_s(h(X_t))-\alpha_{s,t}(P^{\eta_n}_{r,y}))$ \end{enumerate} \label{thmfellerpro} \end{theorem} {\bf Proof} It follows from Corollary \ref {corpen} and Proposition \ref{proppen2} that $\alpha$ is a Feller penalty. Thus from Theorem \ref{ThmTCCP}, and Definition \ref{defTCPP}, $\Pi^{r,y}_{s,t}$ is a time consistent convex dynamic procedure on $\hat {\cal W}^r_t$.\\ With the notations of Proposition \ref{propsup}, let $\tilde h=\sup_{\gamma \in \Theta(\Gamma)^0_t}h^{\gamma}$. From equation (\ref{eqetasj2}), it follows that for all $s$ and all $0 \leq r \leq s$, \begin{equation} \sup_{\gamma \in \Theta(\Gamma)^r_t}h^{\gamma}(s,x)= \sup_{\eta \in \Theta(\Gamma)^s_t}h^{\eta}(s,x) \label{eqtildeh} \end{equation} It follows from Proposition \ref{propsup} and equation (\ref{eqtildeh}) that $\forall r \leq s \leq t$, \begin{eqnarray} \Pi^{r,y}_{s,t}(h(X_t))&=&\sup_{Q^{\gamma}_{r,y} \in {\cal Q}^{\theta}(\Gamma)_{r,y}} (T^{Q^{\gamma}_{r,y}}_s(h(X_t))- \alpha_{s,t}(Q^{\gamma}_{r,y}))\nonumber\\ &=& \sup_{\eta \in \Theta(\Gamma)^r_t} h^{\eta}(s,X_s)=\tilde h(s,X_s) \label{eqtcd8} \end{eqnarray} The last assertion follows from Theorem \ref{ThmTCCP}. \section{Viscosity solution } \label{viscsol} The pocedures are the procedures constructed in the previous Section in case of Diffusions with Levy generator or in case of continuous diffusions. From now on, one assumes furthermore that $\Gamma$ is a multivalued Borel mapping convex and closed valued. \subsection{Viscosity supersolution } One assumes that for all $K $ large enough, $\Gamma_K$ is lower hemicontinuous (cf Definition 16.2 in \cite{AB}, where $\Gamma_K(t,x)=\{(a,b,\mu) \in \Gamma(t,x),\;||a|| \leq K,||b|| \leq K, \int_{\R^n-\{0\}} [||y||^2 1_{||y|| \leq 1}+||y||1_{||y|| >1}]d\mu(y) \leq K\}$ in case of diffusions with Levy generator (and $\Gamma_K(t,x)=\{(a,b) \in \Gamma(t,x),\;||a|| \leq K,||b|| \leq K$ in case of continuous diffusions). Let $(t_0,x_0) \in [0,t[ \times \R^n$. Let $\phi \in {\cal C}^{1,2}_b([0,t] \times \R^n )$ such that \begin{equation} 0=v(t_0,x_0)-\phi(t_0,x_0)=\min(v(t,x)-\phi(t,x)) \label{eqinphi} \end{equation} \begin{lemma} Let $\phi \in {\cal C}^{1,2}_b([0,t] \times \R^n )$. Assume that $M$ satisfies the hypothesis $M_C$ of Definition \ref{deftetacont2}. Let \begin{equation} \tilde K\phi(s,x)(y)=\phi(s,x+y)-\phi(s,x)-\frac{y^*\nabla \phi(s,x)}{1+||y||^2} \label{eqtilk} \end{equation} The function $K\phi(s,x)=\int_{\R^n-\{0\}}\tilde K\phi(s,x)(y)M(s,x, dy)$ is a continuous function of $(s,x)$. \label{lemmakcon} \end{lemma} {\bf Proof} The functions $\tilde K\phi(s,x)$ given by (\ref{eqtilk}) are uniformly bounded. Thus for all $\epsilon>0$, there is $K>0$ such that for all $\mu \in {\cal M}_C$ (Definition \ref{defMc}), \begin{equation} \forall (s,x),\;\int_{||y|| \geq K} \tilde K \phi(s,x)(y)\mu(dy)< \epsilon \label{eq1} \end{equation} It follows from the Taylor formula with integral remainder that $\tilde K \phi(s,x)(y)-\tilde K \phi(s_0,x_0)(y)=y^*(\nabla(\phi)(s,x)-\nabla(\phi)(s_0,x_0))\frac{||y||^2}{1+||y||^2}+ \int_0^1Tr(D^2\phi(s,x+uy)-D^2 \phi(s_0,x_0+uy))yy^*](1-u)du$. Let $K_0>0$. Making use of the uniform continuity of $(s,x) \rightarrow \nabla(f)(s,x)$ and $(s,x) \rightarrow D^2\phi(s,x)$ on compact spaces it follows that for $|s-s_0|<\eta$, $||x-x_0||<\eta$, $||x|| \leq K_0$ and $||x_0|| \leq K_0$, $\int_{||y|| \leq K} |\tilde K\phi(s,x,y)-\tilde K\phi(s_0,x_0,y)| \leq \int_{||y|| \leq K} \epsilon||y||^2\mu(dy)$ for all $\mu \in {\cal M}_C$. Thus $\int_{\R^n-\{0\}} |\tilde K\phi(s,x,y)-\tilde K\phi(s_0,x_0,y)|\mu(dy)$ tends to $0$ uniformly in $\mu \in {\cal M}_C$ when $(s,x)$ tends $(s_0,x_0)$. On the other hand, making use one more time of the Taylor formula and of the continuity hypothesis (Definition \ref{deftetacont2} 2.), it follows that $\int_{\R^n-\{0\}}\tilde K\phi(s_0,x_0,y)M(s,x,dy)$ is a continuous function of $(s,x)$. \hfill $\square $ \begin{theorem} The hypothesis are those of Theorem \ref{thmfellerpro}. Assume also that for $K$ large enough, $\Gamma_K$ is lower hemi continuous. Let $h$ be continuous bounded from below. Let $v=\tilde h$ be a lower semi continuous function such that equations (\ref{eqmeas20}) and (\ref{eqmeas0}) are satisfied. In case of Levy diffusions $v$ is a viscosity supersolution of \begin{equation} \left \{\parbox{12cm}{ \begin{eqnarray} -\partial_u v(u,x)-f(u, x,Dv(u,x),D^2v(u,x),\tilde K v(u,x))&=&0 \nonumber \\ v(t,x)& = & h(x)\nonumber \end{eqnarray} }\right. \label{edpvisc} \end{equation} at each point $(t_0,x_0)$ such that $f(t_0,x_0, D\phi(t_0,x_0), D^2\phi(t_0,x_0), \tilde K \phi(t_0,x_0)))<\infty$ for all $\phi \in {\cal C}_b^{1,2}$ satisfying (\ref{eqinphi}). Here $f(t,x, D \phi(t,x),D^2\phi(t,x),\tilde K \phi(t,x))= \sup_{(a,b,\mu) \in \Gamma(t,x)}[(b^*D\phi(t,x)+\frac{1}{2}Tr(aD^2\phi(t,x))+\int\tilde K \phi(t,x)(y) \mu(dy)-g(t,x,a,b,\mu)]$, and $\tilde K \phi(t,x))$ is given by equation (\ref{eqtilk}) \label{thmsuper0}\\ In case of continuous diffusions, assume that the restriction of $g$ to $\{(u,x,y),\; y \in \Gamma(u,x)\}$ is upper semi-continuous. Then $v$ is a viscosity supersolution of \begin{equation} \left \{\parbox{12cm}{ \begin{eqnarray} -\partial_u v(u,x)-f(u, x,Dv(u,x),D^2v(u,x))&=&0 \nonumber \label{eqedp000}\\ v(t,x)& = & h(x)\nonumber \end{eqnarray} }\right. \end{equation} at each point $(t_0,x_0)$ such that $f(t_0,x_0, D\phi(t_0,x_0),D^2\phi(t_0,x_0))<\infty$ for all $\phi \in {\cal C}_b^{1,2}$ satisfying (\ref{eqinphi}). In this case \begin{equation} f(t,x, z,\gamma)= \sup_{(a,b) \in \Gamma(t,x)}[(b^*z+\frac{1}{2}Tr(a\gamma)-g(t,x,a,b)] \label{eqfjo} \end{equation} \label{thmsuper} \end{theorem} {\bf Proof} The proof follows the proof of Theorem 5.9 of \cite{JBN-PDE}, replacing Ito's formula by the martingale property. From the time consistency for the process $\Pi^{t_0,x_0}_{u,s}$, for all $\delta>0$, \begin{eqnarray} v(t_0,x_0)=\Pi^{t_0,x_0}_{t_0,t}(h(X_t))=\Pi^{t_0,x_0}_{t_0,t_0+\delta}(\Pi^{t_0,x_0}_{t_0 +\delta,t}(h(X_t))) \end{eqnarray} From the definition of $v$ and the inequality $v \geq \phi$, it follows that \begin{equation} \Pi^{t_0,x_0}_{t_0+\delta,t}(h(X_t))=v(t_0+\delta, X_{t_0+\delta}) \geq \phi(t_0+\delta, X_{t_0+\delta}) \label{ineq02} \end{equation} From the definition of $\Pi^{t_0,x_0}_{t_0,t_0+\delta}$, it follows that for all $\theta \in \Theta(\Gamma)$, \begin{equation} v(t_0,x_0) \geq Q^{\theta}_{t_0,x_0}(v(t_0+\delta, X_{t_0+\delta}))-\alpha_{t_0,t_{0+\delta}}( Q^{\theta}_{t_0,x_0}) \label{ineq03} \end{equation} \begin{itemize} \item We give details in case of diffusions with Levy generator. Let $\tilde K\phi(s,x)$ be given by equation (\ref{eqtilk}). The function $\phi$ belongs to ${\cal C}^{1,2}_b$. Thus it follows from the martingale property for $Q^{\theta}_{t_0,x_0}$ as stated in Theorem 1.1 of \cite{St}, from the definition of the penalty (Definition \ref{defpenlev}) and the equality $v(t_0,x_0)=\phi(t_0,x_0)$ that \begin{eqnarray} 0 \geq &Q^{\theta}_{t_0,x_0}[\int_{t_0} ^{t_0+\delta}\{\partial_u \phi(u,X_u)+\frac{1}{2}Tr(aD^2\phi)(u,X_u)+b^*D\phi(u,X_u)]du\nonumber\\ &+Q^{\theta}_{t_0,x_0}[\int_{t_0} ^{t_0+\delta} [ \int_{\R^n-\{0\}}(\tilde K \phi(u,X_u)(y) M(u,X_u, dy))\}du ]\nonumber\\ &+ Q^{\theta}_{t_0,x_0}[(\int_{t_0} ^{t_0+\delta} -g(u,X_u,a(u,X_u),b(u,X_u),M(u,X_u)du ] \label{ineq04} \end{eqnarray} $f(t_0,x_0, D\phi(t_0,x_0),D^2\phi(t_0,x_0),\tilde K \phi(t_0,x_0))= \sup_{(a,b,\mu) \in \Gamma(t_0,x_0)}[b^*D\phi(t_0,x_0)+\frac{1}{2}Tr(aD^2\phi(t_0,x_0))+\int_{\R^n-\{0\}}\tilde K \phi(t_0,x_0)(y) \mu(dy)-g(t_0,x_0,a,b,\mu)]$. Let $(a_0,b_0,\mu_0) \in \Gamma$ such that \begin{eqnarray} &(b_0^*D\phi(t_0,x_0)+\frac{1}{2}Tr(a_0D^2\phi(t_0,x_0))+\int \tilde K\phi(t_0,x_0)d\mu_0-g(t_0,x_0,a_0,b_0,\mu_0)]\nonumber\\ &>f(t_0,x_0, D\phi(t_0,x_0),D^2\phi(t_0,x_0),\tilde K \phi(t_0,x_0))-\epsilon \label{eqcp2} \end{eqnarray} Let $K$ such that $(a_0,b_0,\mu_0) \in \Gamma_K$ and $\Gamma_K$ is lower hemicontinuous. There is then a continuous function $(s,x) \rightarrow a(s,x),b(s,x),M(s,x))=\theta^0(s,x) \in \Gamma_K$ such that $a(s_0,x_0),b(s_0,x_0),M(s_0,x_0)=(a_0,b_0,\mu_0)$. \\ From the regularity properties of $\phi$ and from Lemma \ref{lemmakcon}, one can write (\ref{ineq04}) as \begin{equation} 0 \geq Q^{\theta}_{t_0,x_0}[\int_{t_0} ^{t_0+\delta}(\xi(u,X_u)-g^{\theta^0}(u,X_u))du \label{eqgLevy} \end{equation} where $\xi$ and $g^{\theta^0}$ are continuous bounded. \item In case of continuous diffusions, it follows from equation (\ref{ineq03}) that \begin{equation} 0 \geq Q^{\theta}_{t_0,x_0}[\int_{t_0} ^{t_0+\delta}(\xi(u,X_u)-\overline g(u,X_u))du \label{eqgLevc} \end{equation} where $\xi$ is continuous bounded and $\overline g$ is upper semi-continuous. \end{itemize} For all $\epsilon>0$, there is $\eta>0$, such that for $t_0 \leq u \leq t \leq t_0+\eta$ and $||x_0-x||<\eta$, \begin{eqnarray} \xi(u, x)-\tilde g(u,x) \geq \xi(t_0, x_0)-\tilde g(t_0,x_0)-\epsilon \label{eqsubv} \end{eqnarray} with $\tilde g$ equal $g^{\theta^0}$ or $\overline g$. In case of Levy generators, let $K_0=||\xi||_{\infty}+||g^{\theta^0}||_{\infty}$. In case of continuous diffusions, from hypothesis $H_g$, $|\overline g(u,x)|\leq 1+||x||^m$. Let $K_0=||\xi||_{\infty}+1 +(Q^{\theta_0}_{t_0,x_0}(\sup_{t_0 \leq u \leq \eta}||X_u||^{2m})^{\frac{1}{2}}$. From Lemma \ref{lemmaLM} in case of Levy generators and from Proposition 2.3 of \cite{JBN-PDE} in case of continuous diffusions, applied with the probability measure $Q^{\theta_0}_{t_0,x_0}$ there is $0<\delta < \eta$ such that \begin{equation} Q^{\theta_0}_{t_0,x_0}(A)<\frac{\epsilon}{K_0}\;\; with \;\;A=\{\omega \;|\; \sup_{t_0 \leq u \leq t_0+ \delta}||X_u-x_0||>\eta\;\} \label{eqcpC} \end{equation} Dividing the inequality (\ref{ineq04}) by $\delta$, making use of (\ref{eqcp2}), (\ref{eqsubv}), (\ref{eqcpC}) and of Cauchy Schwarz inequality, it follows that $$0>f(t_0,x_0, D\phi(t_0,x_0,D^2\phi(t_0,x_0)\tilde K \phi(t_0,x_0))-3\epsilon$$. \hfill $\square $ \begin{remark} A different approach to viscosity solutions of fully non linear PDE is proposed in \cite{STZ2}. This approach is based on the existence of solutions to BSDE. In \cite{STZ2} the PDE is $$-\partial_u v(u,x)-f(u, x,v(u,x),Dv(u,x),D^2v(u,x)) $$ with $f(t,x, y,z,\gamma)= \sup_a[\frac{1}{2}Tr(a\gamma)-g(t,x,y,z,a)]$, $f$ is allowed to depend on $y$. However unlike for equation (\ref{eqfjo}), the assumptions on function $g$ (Assumption 5.1 of \cite{STZ2}) are quite restrictive: the domain of $g(t,x,.)$ is independent of $t$ and $x$, the function $g$ is uniformly continuous in $t$, Lipschitz in $z$. \end{remark} \subsection{Viscosity subsolution} From now on we add a new hypothesis $H'$. In case of continuous diffusions we assume that the multivalued Borel mapping $\Gamma$ has linear growth, i.e. There is a constant $K>0$ such that $ \forall (a,b) \in \Gamma(s,x),\; ||a||,||b|| \leq K(1+||x||)$. As in \cite{JBN-PDE} Lemma 5.11, it follows from \cite{Kry} II 5 Corollary 10 that for all $q\geq 1$, $C$, and $t>0$, there is a constant $K_1$ such that for all $y$ such that $||y|| \leq C$ and all bounded $\gamma \in \Theta(\Gamma)^r_t$, \begin{equation} E_{Q^{\gamma}_{r,y}}(\sup_{s \leq u \leq t }(||X_t-X_{u}||^{2q}) \leq K_1 (t-s)^q \label{eqKry01} \end{equation} and \begin{equation} \forall \epsilon>0,\forall \eta>0,\exists h>0 \forall r>0, \forall y, ||y|| \leq C,\; \forall \gamma\in \Theta(\Gamma)^r_t, \; Q^{\gamma}_{r,y}[\sup_{r \leq v \leq r+h}||X_v-y||>\eta)< \epsilon \label{equnif} \end{equation} Denote $v^*$ the upper semi continuous envelope of $v$, $$v^*(s,x)=\limsup_{(s',x')\rightarrow (s,x)} v(s',x')$$ Let $\phi \in {\cal C}^{1,2}_{b}$ such that \begin{equation} 0=v^*(t_0,x_0)-\phi(t_0,x_0)=\sup_{(s,x)}(v^*(s,x)-\phi(s,x)) \label{equsc} \end{equation} \begin{theorem} Let $h$ be continuous bounded from below. Assume that \begin{enumerate} \item In case of continuous diffusions, $g$ satisfies hypothesis $H_g$ and $\Gamma$ is closed convex with linear growth. \item In case of diffusions with Levy generator, $\Gamma=\Gamma_K$ for some $K>0$. \end{enumerate} Assume that $f$ is upper semi continuous. Let $v=\tilde h$ be the lower semi continuous function as in Theorem \ref{thmfellerpro}. Then $v^*$ is a viscosity subsolution of (\ref{eqedp00}) in case of continuous diffusions and of (\ref{edpvisc}) in case of diffusions with Levy generator \label{thmsub} \end{theorem} $\tilde K \phi(u,x)$ belongs to the vector space ${\cal C}(\R^n)$ of continuous functions on $\R^n$. Let ${\cal M}$ be the vector space of measures generated by ${\cal M}_C$. The topology on $E$ is $\sigma(E,{\cal M})$.\\ {\bf Proof} It follows from the first part of the proof of Lemma \ref{lemmakcon} that $\tilde K \phi$ is continuous from $[0,T] \times \R^n$ to $E$. The functions $\phi_u$,$D\phi$, $D^2\phi$ are continuous and $f$ is upper semi-continuous thus for all $n>0$ there is $\eta_n>0$ such that for $t_0 \leq u \leq t_0+\eta_n$, and $||x-x_0||<\eta_n$, $\phi_u(u,x)+f(u,x, D\phi(u,x), D^2\phi(u,x),\tilde K \phi(u,x) \leq$\\ $\phi_u(t_0,x_0)+f(t_0,x_0, D\phi(t_0,x_0), D^2\phi(t_0,x_0),\tilde K \phi(t_0,x_0)+\frac{1}{n}$. From equation (\ref{equnif}) in case of continuous diffusions and inequality (\ref{eqLM}) in case of diffusions with Levy generator, there is $\alpha_n>0$ such that for all $|u-t_0|\leq 1$ and $||y-x_0|| \leq 1$ for all $\gamma \in \Theta(\Gamma)^u_t$, $${Q^{\gamma}_{u,y}}(A_n)<\frac{1}{n}\;\; \text{with} \;\;A_n=\{\omega \;|\; \sup_{u \leq u' \leq u+ \alpha_n}||X_{u'}-y||>\frac{\eta_n}{2}\;\}$$ Let $\delta_n=inf(\eta_n,\alpha_n)$. For all $n>0$ choose $(t_n,x_n)$ such that $|t_n-t_0|<\frac{\delta_n}{2}$, $||x_n-x_0||<\frac{\eta_n}{2}$ and $\phi(t_n,x_n)- v(t_n,x_n) \leq \frac{\delta_n}{n}$. For all $n>0$, there is a process $\gamma_n \in \Theta(\Gamma)^{t_n}_t$, such that \begin{equation} v(t_n,x_n) \leq E_{Q^{\gamma_n}_{t_n,x_n}}(h(X_t))-\alpha_{t_n,t}(Q^{\gamma_n}_{t_n,x_n})+\frac{1}{n}. \label{eq24} \end{equation} From the cocycle condition for the penalty associated to the probability measure $Q^{\gamma_n}_{t_n,x_n}$, the definition of $\Pi^{t_n,x_n}_{t_n+\delta_n,t}(h(X_t)$ and $\Pi^{t_n,x_n}_{t_n+\delta_n,t}(h(X_t)=v(t_n + \delta_n,X_{t_n + \delta_n}) \leq \phi(t_n + \delta_n,X_{t_n + \delta_n})$ \begin{equation} v(t_n,x_n) \leq E_{Q^{\gamma_n}_{t_n,x_n}}[\phi(t_n+\delta_n,X_{t_n+\delta_n})-\int_{t_n}^{t_n+\delta_n}g (u,X_u,\mu_n(u,\omega)du]+\frac{1}{n} \label{eq25} \end{equation} We apply the martingale property for $Q^{\gamma_n}_{t_n,x_n}$ and we make use of the definition of $f$. We can then finish the proof as the proof of the viscosity supersolution. \subsection{Viscosity solution and uniqueness} \label{subsecU} The following Theorem results from Theorems \ref{thmsuper} and \ref{thmsub} \begin{theorem} Assume that all the previous hypothesis are satisfied. Assume furthermore that $h $ is bounded from below and $\alpha$ H\"older-continuous for some $\alpha>0$ in case of continuous diffusions and $h$ is Lipschitz in case of diffusions with Levy generator. Let $v=\tilde h$ be a lower semi continuous function such that equations (\ref{eqmeas20}) and (\ref{eqmeas0}) are satisfied. Assume that the PDE (\ref{edpvisc}) satisfies the comparison principle for functions bounded on compact spaces. Then the function $v$ is continuous. It is the unique viscosity solution of (\ref{edpvisc}). \end{theorem} {\bf Proof} As in \cite{JBN-PDE} we prove first that $v$ is continuous at $(t,x)$ for all $x$. Making use of the equation (\ref{eqKry01}) in case of continuous diffusions and of Lemma \ref{lemmaLM} for diffusions with levy generator, we prove as in \cite{JBN-PDE} that $-v(s,x)-h(x)$ tends to $0$ uniformly in $x$ when $s$ tends to $t$. The result follows then from Thm \ref{thmsuper} and \ref{thmsub} and from Proposition 5.5 of \cite{JBN-PDE}. For comparison results for non linear second order PDE we refer to \cite{GIL,FSo,Da}.\\
\section*{Introduction} Our original motivation for this work was to generalize some of the results obtained by Herzog, Takayama, and Terai in \cite{HTT}. They proved that many properties of a monomial ideal pass to its radical. It is well known that every monomial ideal $I$ in the polynomial ring $S=K[x_1,\ldots,x_n]$ over a field $K$ is a positively $\tb$-determined $S$-module for an appropriate $\tb\in \NN^n$ as it was defined by Miller in \cite{Mill}. Thus, a natural way to generalize the results of \cite{HTT} is to consider positively determined modules instead of monomial ideals. We show that one may define a functor $\rr^\ast$ from the category $\Modtb$ of positively $\tb$-determined modules to the category $\Modone$ of squarefree $S$-modules which plays the role of passing from a monomial ideal to its radical. As it was shown in \cite{BF}, to any ordering preserving map $\qb:\NN^n\to \NN^n$, one may associate a functor $\qb^\ast$ from the category of $\NN^n$-graded $S$-modules to itself which is defined as follows. For any $M\in \ModN$ and $\ab\in \NN^n,$ $(\qb^\ast M)_{\ab}=M_{q(\ab)}$ and the $S$-module structure of $\qb^\ast M$ is given by the multiplication $(q^\ast M)_{\bb}\stackrel{\cdot\xb^\ab}{\rightarrow}(q^\ast M)_{\bb+\ab}$ that maps every homogeneous element $y\in (q^\ast M)_{\bb}$ to $\xb^{q(\ab+\bb)-q(\bb)}y.$ If $f:M\to N$ is a graded morphism of $\NN^n$-graded $S$-modules, then for every $\ab\in \NN^n,$ the $\ab$-degree component of $q^\ast f: q^\ast M \to q^\ast N$ is $f_{q(\ab)}.$ In \cite{BF} it is shown that $q^\ast$ is an exact functor. In Section~\ref{radicalsection}, we have considered the following map. For $\tb\in \NN^n$ with $\tb\geq \One$ where $\One=(1,\ldots,1), $ we define $\rr: \NN^n \to \NN^n$ by $\rr(\ab)=(r_i(a_i))_{1\leq i\leq n}$ where \[ r_i(a_i)=\left\{ \begin{array}{ll} t_i, & \text{ if } a_i>0,\\ 0, & \text{ otherwise. } \end{array}\right. \] This is an ordering preserving map which induces a functor $\rr^\ast$ which depends on $\tb$ from the category $\ModN$ to itself. We showed in Section~\ref{radicalsection} that this functor transports the category $\Modtb$ into the category of squrefree modules and, moreover, for any monomial ideal $I\subset S,$ we have $\rr^\ast I\cong \sqrt{I}$ as $S$-modules. As it was explained in \cite{HTT}, the Betti numbers do not increase when one passes from a monomial ideal to its radical. We show in Theorem~\ref{bettiradical} that passing from a positively $\tb$-determined module to its ''radical'' module has a similar behavior. In particular, one obtains $\depth M\leq \depth\rr^\ast M$ for any $M\in \Modtb$. Unlike the monomial case, for a positively $\tb$-determined module $M,$ we show in Corollary~\ref{dim}, that one has only the inequality $\dim \rr^\ast M\leq \dim M.$ Easy examples show that the inequality may be strict. By using the inequalities between depth and Krull dimension, we show in Theorem~\ref{CM} that the (sequentailly) Cohen-Macaulay property of $M$ passes to the ''radical'' of $M$ for any postively $\tb$-determined module $M$with $\rr^\ast M\neq 0.$ In Section~\ref{extsection} we study the connection between the functor $\rr^\ast$ and $\Ext$. The main result of the section is Theorem~\ref{ext} which states that for every module $M\in \Modtb$ there exists a natural isomorphism $ \grExt^p_S(M, \omega_S)_{\geq \Zero} \cong \grExt^p_S(\rr^\ast(M),\omega_S)$ for all $p,$ where $\omega_S$ is the canonical module of $S.$ In particular, for any Cohen-Macaulay ideal $I\subset S$ such that $S/I$ is a positively $\tb$-determined module, it follows that the canonical module of $S/\sqrt{I}$ is isomorphic to the positive part of $\omega_{S/I}.$ From Theorem~\ref{ext}, under an additional condition, it follows that if $M\in \Modtb$ is a generalized Cohen-Macaulay (Buchsbaum) module, then $\rr^\ast M$ shares the same property. Finally, in the last two sections we show how our ''radical'' functor is connected to the Alexander duality (Proposition~\ref{prop:r_and_A}) and Auslander-Reiten translate functor (Proposition~\ref{ART}). \section*{Acknowledgment} A part of this research was done when the second author visited the Faculty of Mathematics and Informatics of Ovidius University. He is deeply grateful for warm hospitality. \section{The $\rr^\ast$ functor and first applications} \label{radicalsection} In this section we define the $\rr^\ast$ functor on the category $\ModN$ of the $\NN^n$-graded $S$-modules where $S=K[x_1,\ldots,x_n]$ is the polynomial ring in $n$ variables over a field $K.$ We first recall the basic notions and set the notation. Let $\NN$ be the set of non-negative integers. For $\ab=(a_1,\ldots,a_n)\in \NN^n$ we set $\xb^\ab=x_1^{a_1}\cdots x_n^{a_n}$ and call $\ab$ the degree of the monomial $\xb^\ab$. $\nu_i(u)$ denotes the exponent of variable $x_i$ in the monomial $u\in S$. Let $\leq$ be the partial order on $\ZZ^n$ which is defined as follows. If $\ab=(a_1,\ldots,a_n),\bb=(b_1,\ldots,b_n)\in\ZZ^n$, then $\ab\leq \bb$ if $a_i\leq b_i$ for $1\leq i\leq n.$ Of course, this order induces a partial order on $\NN^n.$ Let $\tb\in \NN^n$ with $\tb\geq \One,$ where $\One=(1,\ldots,1).$ According to \cite{Mill}, a $\ZZ^n$-graded $S$-module $M$ is called {\em positively $\tb$-determined} if it is finitely generated, $\NN^n$-graded, and if the multiplication map $M_{\ab}\stackrel{x_i}{\rightarrow}{M_{\ab+\eb_i}}$ is an isomorphism of $K$-vector spaces whenever $a_i \ge t_i$. Here, $\eb_i$ is the vector of $\ZZ^n$ with its $i$-th component equal to $1$ and all the others equal to $0.$ A monomial ideal $I$ is positively $\tb$-determined if and only if it is generated by some elements $x^\ab$ with $\Zero \le \ab \le \tb$. Every finitely generated $\NN^n$-graded $S$-module is positively $\tb$-determined for some $\tb \gg \One$. In particular, for any 2 monomial ideals $I,J$ with $J \supseteq I$, $J/I$ is positively $\tb$-determined for some $\tb \gg \One$. Let $\Modtb$ be the full subcategory of $\ModZ$ consisting of positively $\tb$-determined $S$-modules. According to \cite{BF}, with any order preserving map $q:\ZZ^n\to \ZZ^n,$ one may associate a functor $q^{\ast}: \ModZ \to \ModZ$. Since we are concerned only with $\NN^n$-graded modules, that is, $\ZZ^n$-graded modules whose components of degree $\ab\in \ZZ^n\setminus \NN^n$ are all zero, we may consider the map $q:\NN^n\to \NN^n.$ $q^\ast$ acts on modules and morphisms as follows. For a $\ZZ^n$-graded $S$-module $M,$ the $\ab$-degree component of $q^\ast M$ is $(q^\ast M)_{\ab}=M_{q(\ab)}.$ The multiplication which gives the $S$-module structure of $q^\ast M$ is the following. For a monomial $\xb^\ab\in S,$ the map $(q^\ast M)_{\bb}\stackrel{\cdot\xb^\ab}{\rightarrow}(q^\ast M)_{\bb+\ab}$ maps every homogeneous element $y\in (q^\ast M)_{\bb}$ to $\xb^{q(\ab+\bb)-q(\bb)}y.$ We describe now the action of $q^\ast$ on the morphisms of the category $\ModZ$ following \cite{BF}. If $f:M\to N$ is a graded morphism of $\ZZ^n$-graded $S$-modules, then for every $\ab\in \ZZ^n,$ the $\ab$-degree component of $q^\ast f: q^\ast M \to q^\ast N$ is $f_{q(\ab)}.$ In \cite{BF} it is shown that $q^\ast$ is an exact functor. We consider now the following order preserving map. Let $\tb \in \NN^n, \tb\geq \One,$ and $\rr: \NN^n \to \NN^n$ given by $\rr(\ab)=(r_i(a_i))_{1\leq i\leq n}$ where \[ r_i(a_i)=\left\{ \begin{array}{ll} t_i, & \text{ if } a_i>0,\\ 0, & \text{ otherwise. } \end{array}\right. \] It is easily seen that $\rr$ is an order preserving map, hence we may consider the functor $\rr^\ast: \ModN\to \ModN$ associated with $\rr.$ \begin{Proposition}\label{sqfree} Let $M$ be a positively $\tb$-determined $\NN^n$-graded $S$-module. Then $\rr^\ast M$ is a positively $\One$-determined module, that is, $\rr^\ast M$ is a squarefree $S$-module. \end{Proposition} \begin{proof} It is enough to show that, for any $\ab\in \NN^n$ and $i\in \supp(\ab),$ the multiplication map \[ (\rr^\ast M)_{\ab} \stackrel{\cdot x_i}{\rightarrow}(\rr^\ast M)_{\ab+\eb_i} \] is an isomorphism of $K$-vector spaces. But this is almost obvious, since $\xb^{e_i}\cdot u=\xb^{\rr(\ab+\eb_i)-\rr(\ab)}u=u.$ The last equality is true since $\supp(\xb^{\rr(\ab+\eb_i)-\rr(\ab)})=\emptyset$ for any $i\in \supp(\ab)$. Therefore, the multiplication by $x_i$ is the identity map, hence it is an isomorphism of vector spaces. \end{proof} \begin{Proposition}\label{classicradical} Let $\tb\in \NN^n$ with $\tb\geq \One,$ and let $I\subset S$ be a monomial ideal which is positively $\tb$-determined, that is, for every $u\in G(I),$ $\deg(u)\leq \tb.$ Then $\rr^\ast I\cong \sqrt{I}.$ \end{Proposition} \begin{proof} Firstly, we claim that $\xb^\ab\in \sqrt{I}$ if and only if $\xb^{\rr(\ab)}\in I.$ Let us prove this claim. If $\xb^{\ab}\in \sqrt{I},$ then there exists $k\geq 1$ such that $\xb^{k\ab}\in I.$ Obviously, we may choose $k$ such that $ka_i\geq t_i$ for all $a_i>0.$ Then there exists $\xb^\bb\in G(I)$ such that $\xb^\bb\ |\ \xb^{k\ab},$ which implies that $\bb\leq k\ab.$ Since $I$ is positively $\tb$-determined, we also have $\bb\leq \tb.$ It then follows that $\bb\leq \rr(k\ab)=\rr(\ab)$ which implies that $\xb^\bb\ |\ \xb^{\rr(\ab)}$ and, therefore, $\xb^{\rr(\ab)}\in I.$ Conversely, let $\xb^{\rr(\ab)}\in I.$ We obviously may find $k\geq 1$ such that $k\ab\geq \rr(\ab),$ hence $\xb^{k\ab}\in I$ and, therefore, $\xb^{\ab}\in \sqrt{I},$ which ends the proof of our claim. Let $f:\sqrt{I}\to \rr^\ast I$ be the map given by $f=\oplus_{\ab}f_{\ab}$ where $f_{\ab}:(\sqrt{I})_\ab\to (\rr^\ast I)_{\ab}$ is defined by $f_{\ab}(\xb^\ab)=\xb^{\rr(\ab)}.$ The map $f$ is obviously a graded isomorphism of $K$-vector spaces. We show that $f$ is an $S$-module isomorphism. Indeed, for any $\ab, \bb,$ \[ f(\xb^\bb \cdot \xb^\ab)=\xb^{\rr(\ab+\bb)}= \xb^{\rr(\ab+\bb)-\rr(\ab)}\xb^{\rr(\ab)}= \xb^\bb\cdot f(\xb^\ab). \] \end{proof} In order to state the first main result, we need a preparatory lemma. Before stating it, let us set some more notation. For $\ab\in \NN^n$, let $S(-\ab)$ be the graded free $S$-module whose all graded components are obtained from those of $S$ by shifting with the vector $\ab,$ and let $\sqrt{\ab}$ be the following vector of $\NN^n:$ \[ (\sqrt{\ab})_i=\left\{ \begin{array}{ll} 1, & \text{ if } a_i>0,\\ 0, & \text{ if } a_i=0. \end{array}\right. \] \begin{Lemma}\label{shift} Let $\tb\in \NN^n$ with $ \tb\geq \One.$ Then $r^\ast (S(-\ab))\cong S(-\sqrt{\ab})$ for every $\ab\in \NN^n$ with $\ab\leq \tb.$ \end{Lemma} \begin{proof} We obviously have the following isomorphisms: \[ \rr^\ast(S(-\ab))\cong \rr^\ast(\xb^\ab)\cong (\xb^{\sqrt{\ab}})\cong S(-\sqrt{\ab}). \] \end{proof} In the sequel we will always assume that $\rr^\ast M\neq 0$. Note that $\rr^\ast M=0$ if and only if $M_{\ab}=0$ for all $\ab\in \NN^n$ such that $a_i\in\{0,t_i\}$ for $1\leq i\leq n.$ The following theorem shows that the graded Betti numbers go down when passing from the module to its radical. In particular, we may derive inequalities for the corresponding depths. \begin{Theorem}\label{bettiradical} Let $\tb\in \NN^n,$ $\tb\geq \One$, and let $M$ be a positively $\tb$-determined $\NN^n$-graded $S$-module. Then \[ \beta_{i,\ab}(M)\geq \beta_{i,\sqrt{\ab}}(\rr^\ast M) \] for all $i$ and $\ab.$ In particular, the following inequality holds: \[ \depth M\leq \depth \rr^\ast M. \] \end{Theorem} \begin{proof} Let \[ \FF_{\bullet}: \hspace{0.8cm} 0\to \bigoplus_{\ab}S(-\ab)^{\beta_{p,\ab}}\to \cdots \to \bigoplus_{\ab}S(-\ab)^{\beta_{1,\ab}} \to \bigoplus_{\ab}S(-\ab)^{\beta_{0,\ab}}\to M\to 0 \] be a minimal free resolution of $M$ over $S.$ Since $M$ is positively $\tb$-determined, by \cite[Proposition 2.5]{Mill}, it follows that all the shifts in the above resolution are $\leq \tb.$ We apply $\rr^\ast$ to $\FF_{\bullet}$. By the exactness of $\rr^\ast,$ we get a free $S$-resolution of $\rr^\ast M,$ possibly non-minimal. Therefore, we get the inequalities between the graded Betti numbers of $M$ and, respectively, $\rr^\ast M.$ These inequalities imply that $\projdim_S M\geq \projdim_S( \rr^\ast M)$ and, by using Auslander-Buchsbaum formula, we get the inequalities between depths. \end{proof} \begin{Remark}\label{resol}{\em By the above proof it follows that if \[ \FF_{\bullet}: 0\to \bigoplus_{\ab}S(-\ab)^{\beta_{p,\ab}}\to \cdots \to \bigoplus_{\ab}S(-\ab)^{\beta_{1,\ab}} \to \bigoplus_{\ab}S(-\ab)^{\beta_{0,\ab}}\to M\to 0 \] is a minimal $\ZZ^n$-graded free resolution of $M,$ then \[ \rr^\ast\FF_{\bullet}: 0\to \bigoplus_{\ab}S(-\sqrt{\ab})^{\beta_{p,\ab}}\to \cdots \to \bigoplus_{\ab}S(-\sqrt{\ab})^{\beta_{1,\ab}} \to \bigoplus_{\ab}S(-\sqrt{\ab})^{\beta_{0,\ab}}\to \rr^\ast M\to 0 \] is a free resolution of $\rr^\ast M$. Moreover, if the map $\partial_i: \bigoplus_{\ab}S(-\ab)^{\beta_{i,\ab}} \to \bigoplus_{\ab}S(-\ab)^{\beta_{i-1,\ab}}$ is given by the matrix $(\xb^{\ab_j-\bb_k})_{j,k}$, then $\rr^\ast \partial_i: \bigoplus_{\ab}S(-\sqrt{\ab})^{\beta_{i,\ab}} \to \bigoplus_{\ab}S(-\sqrt{\ab})^{\beta_{i-1,\ab}}$ is given by $(\xb^{\sqrt{\ab_j}-\sqrt{\bb_k}})_{j,k}$. } \end{Remark} In the following corollary we derive some consequences of Theorem~\ref{bettiradical}. To begin with, let $I\subset S$ be a monomial ideal. Then $I$ is a positively $\tb$-determined $\NN^n$-graded $S$-module if we choose, for instance, $\tb=(t_1,\ldots,t_n)$ where $t_i=\max\{\nu_i(u)\ |\ u\in G(I)\}$. Here $G(I)$ denotes the minimal system of monomial generators of the ideal $I.$ Proposition~\ref{classicradical} says that $\rr^\ast I$ is actually $\sqrt{I}.$ Therefore, from Theorem~\ref{bettiradical}, we obtain as a consequence the following corollary which extends results of \cite{HTT}. \begin{Corollary}\label{extensionHHT} Let $I\subset J\subset S$ be monomial ideals with $\sqrt{I}\neq \sqrt{J}$. Then $$\beta^S_{i,\ab}(J/I)\geq \beta^S_{i,\sqrt{\ab}}(\sqrt{J}/\sqrt{I})$$ for all $i\geq 0$ and $\ab\in \NN^n.$ Consequently, $$\depth_S(\sqrt{J}/\sqrt{I})\geq \depth_S(J/I).$$ \end{Corollary} In the following we study the relationship between the Krull dimension of $M$ and $\rr^\ast M.$ We first introduce the following notation. For $\ab\in \NN^n,$ $\supp(\ab)=\{i: a_i>0\}$ and $\supp^\tb(\ab)=\{i: a_i\geq t_i\}.$ We use the following convention. For $\ab, \bb \in \ZZ^n$, let $\ab \cdot \bb$ denote the vector whose $i$-th component is $a_ib_i$. \begin{Proposition}\label{Ass} Let $M$ be a positively $\tb$-determined $S$-module where $\tb\in \NN^n, \tb\geq \One.$ Then $\Ass(\rr^\ast M)\subset \Ass(M).$ \end{Proposition} \begin{proof} Let $F\subset [n]$ and $P=P_F:=(x_i : i\notin F)$ be an associated prime of $\rr^\ast(M)$. Then, by \cite[Lemma 2.2]{Y}, there exists $0\neq u\in (\rr^\ast M)_{\eb_F}$ such that $x_i u=0$ for all $i\notin F,$ where $\eb_F:=\sum_{i\in F}\eb_i.$ This means that there exists $0\neq u\in M_{\tb\cdot \eb_F}$ such that \[ \xb^{\rr(\eb_{F\cup\{i\}})-\rr(\eb_i)} u=\xb^{\tb\cdot \eb_i}u=x_i^{t_i}u=0 \] for all $i\notin F.$ Then we may choose a maximal monomial (with respect to divisibility) $w\in K[\{x_i : i\notin F\}]$ such that $wu\neq 0.$ We claim that $P_F=\ann(wu)$, which will end the proof. If $i\notin F,$ then $x_i(wu)=0$ by the choice of $w,$ hence $P_F\subset\ann(wu).$ Let now $v$ be a monomial in $\ann(wu)$, that is, $v(wu)=0$. Clearly, for every monomial $w^\prime\in K[\{x_i: i\in F\}]$, we have $w^\prime wu\neq 0$ since $\supp(w^\prime)\subset \supp^\tb(wu)$ and $M$ is positively $\tb$-determined. This implies that there exists $i\notin F$ such that $x_i$ divides $v$, thus $v\in P_F.$ \end{proof} \begin{Corollary}\label{dim} Let $M$ be a positively $\tb$-determined module. Then $\dim M\geq \dim \rr^\ast M.$ \end{Corollary} Note that the inequality $\dim \rr^\ast M\leq \dim M$ may be strict as the following example shows. On the other hand, we have $\dim \rr^\ast M = \dim M$ if and only if there exists $\ab\in \NN^n$ such that $\#\supp^\tb(\ab)=\dim M$ and $M_{\rr(\ab)}\neq 0$. \begin{Example}{\em Let $I=(a^4d^4,a^2b^3,b^3c^2,b^3d)$ and $J=(a^3d^3,a^3b,b^2)$, $I,J\subset K[a,b,c,d]$. One may easily check that $\dim(\sqrt{J}/\sqrt{I})=1 < \dim(J/I)=2.$ } \end{Example} Let us recall that a finitely generated $S$-module is called {\em equidimensional} if all its minimal primes have the same codimension. As an immediate consequence of Proposition~\ref{Ass} we get also the following \begin{Corollary}\label{equidim} Let $M$ be a positively $\tb$-determined module such that $\dim M=\dim \rr^\ast M.$ If $M$ is equidimensional, then $\rr^\ast M$ is equidimensional, too. \end{Corollary} The following example shows that the implication of the above corollary is no longer true if $\dim M > \dim \rr^\ast M$. \begin{Example}{\em Let $P,P_1,P_2\subset S,$ $P=(x_1), P_2=(x_1,x_2)$, $P_2=(x_1,x_3,x_4),$ and $M=(S/P)(-(1,0,\ldots,0))\oplus (S/P_1)\oplus (S/P_2)$. Then $M$ is positively $\tb$-determined, where $\tb=(2,1,\ldots,1),$ and equidimensional. We have $\rr^\ast M=(S/P_1)\oplus (S/P_2)$, thus $\rr^\ast M$ is not equidimensional. But, of course, $\dim M > \dim \rr^\ast M.$ } \end{Example} In \cite{HTT} it is shown that if $S/I$ is (sequentially) Cohen-Macaulay, then $S/\sqrt{I}$ shares the same property. We are going to extend this result to any positively determined $\NN^n$-graded module. \begin{Theorem}\label{CM} Let $M$ be a positively $\tb$-determined $S$-module with $\rr^\ast M \neq 0$ where $\tb\in \NN^n, \tb\geq \One.$ \begin{itemize} \item [(i)] If $M$ is Cohen-Macaulay, then $\rr^\ast M$ is Cohen-Macaulay and $\dim \rr^\ast M=\dim M.$ \item [(ii)] If $M$ is sequentially Cohen-Macaulay, then $\rr^\ast M$ is sequentially Cohen-Macaulay. \end{itemize} \end{Theorem} \begin{proof} (i). By Theorem~\ref{bettiradical} and Corollary~\ref{dim}, we get the following inequalities: \[ \depth M\leq \depth \rr^\ast M\leq \dim \rr^\ast M\leq \dim M. \] Since $M$ is Cohen-Macaulay, we get the desired conclusions. (ii). As $M$ is sequentially Cohen-Macaulay, there exists a finite filtration \[ 0=M_0\subset M_1\subset \cdots \subset M_r=M \] by graded submodules of $M$ such that each quotient $M_i/M_{i-1}$ is Cohen-Macaulay and \[ \dim M_1/M_0 < \dim M_2/M_1 <\cdots <\dim M_r/M_{r-1}. \] This filtration induces the following filtration of $\rr^\ast M,$ \[ 0=\rr^\ast M_0\subset \rr^\ast M_1\subset \cdots \subset \rr^\ast M_r=\rr^\ast M. \] By (i), each factor in this filtration is either $0$ or a Cohen-Macaulay module with $\dim \rr^\ast M_i/\rr^\ast M_{i-1}=\dim M_i/M_{i-1}.$ By skipping the redundant factors in the above filtration we get the desired filtration for $\rr^\ast M.$ Therefore, (ii) follows. \end{proof} We may now derive the following corollary which extends some results of \cite{HTT}. \begin{Corollary} \label{CMextend} Let $I\subset J\subset S$ be monomial ideals such that $\sqrt I\neq \sqrt J$. Then: \item [(i)] If $J/I$ is Cohen-Macaulay, then $\sqrt J/\sqrt I $ is Cohen-Macaulay and, moreover, $$\dim J/I=\dim \sqrt J/\sqrt I.$$ \item [(ii)] If $J/I$ is sequentially Cohen-Macaulay, then $\sqrt J/\sqrt I$ is sequentially Cohen-Macau\-lay. \end{Corollary} \section{The functor $\rr^\ast$ and $\Ext$} \label{extsection} For $M \in \ModZ$ and $\ab \in \ZZ^n$, $M(\ab)$ denotes the $\ZZ^n$-graded $S$-module such that $M = M(\ab)$ as underlying $S$-modules and the degree is given by the formula $M(\ab)_\bb = M_{\ab + \bb}$. Following the usual convention, for $M, N \in \ModZ$, we set $\grHom_S(M,N) := \bigoplus_{\ab\in \ZZ^n}\Hom_{\ModZ}(M,N(\ab))$. Note that if $M$ is finitely generated, $\grHom_S (M,N) = \Hom_S(M,N)$. Let $\grExt^i_S(-,N)$ (resp. $\grExt^i_S(M,-)$) be the $i$-th right derived functor of $\grHom_S(-,N)$ (resp. $\grHom_S(M, -)$). In this section, we will study the relation between $\rr^\ast$ and $\grExt$-functor. For this purpose, we need the following three functors. Recall that degree-shifting induces an endofunctor of $\ModZ$. For $\ab \in \ZZ^n$, let $\sigma_\ab$ denote the functor given by shifting degree by $\ab$. Thus we have $\sigma_\ab(M) = M(-\ab)$ for all $M \in \ModZ$. For $M \in \ModZ$ and $\ab \in \ZZ^n$, the truncated module $\tau_{ \ab}(M) := \bigoplus_{\bb \ge \ab} M_\bb$ is again an object of $\ModZ$, and any morphism $M \to N$ in $\ModZ$ induces the one $f|_{\tau_{ \ab}(M)}:\tau_{ \ab}(M) \to \tau_{ \ab}(N)$. Thus we have the functor $\tau_{ \ab}: \ModZ \to \ModZ$. Let $\ss : \NN^n \to \ZZ^n$ be the function defined by $\ss(\ab) = (s_i(a_i))_{1 \le i \le n}$ where $$ s_i(a_i) = \begin{cases} t_i &\text{if $a_i \ge 1$,}\\ t_i - 1 &\text{otherwise.} \end{cases} $$ The induced functor $\ss^\ast: \ModZ \to \ModN$, if restricted to $\Modtb$, is an endofunctor of $\Modtb$. For $M \in \Modtb$ and $\ab \in \ZZ^n$ with $\ab \ge \Zero$, the multiplication $$ x^{\ab \cdot \tb - \rr(\ab)}: M_{\rr(\ab)} \to M_{\ab \cdot \tb} $$ is a $K$-linear isomorphism since $\supp(\ab\cdot\tb-\rr(\ab))\subset \supp^\tb\rr(\ab)$. Let $\phi^M_\ab$ denote this map. Now we are ready to define the two natural transformations $\Phi^\rr: \id_{\Modtb} \Longrightarrow \rr^\ast$ between endofunctors of $\Modtb$ and $\Psi: \ss^\ast \Longrightarrow \sigma_{\One - \tb}$ between those of $\ModN$. For $M \in \Modtb$, let $\Phi_M : M \to \rr^\ast(M)$ be the map defined as follows; for a homogeneous $u \in M_\ab$ with $\ab \ge \Zero$, $$ \Phi_M(u) := {(\phi^M_{\ab})}^{-1}(x^{\ab \cdot (\tb - \One)}u) \in \rr^\ast(M)_\ab. $$ For $\ab \in \NN^n$, it is easy to verify that $\ab + \tb - \One \ge \ss(\ab)$. For $M \in \ModN$, we define the map $\Psi_M: \ss^\ast(M) \to \sigma_{\One - \tb}(M)$ as follows; for $\ab \in \NN^n$ and a homogeneous $u \in \ss^\ast(M)_\ab$ with $\ab \ge \Zero$, $$ \Psi_M(u) = x^{\ab + \tb - \One - \ss(\ab)} \cdot u \in \sigma_{\One - \tb}(M)_\ab. $$ \begin{Lemma}\label{lem:nat} The following statements hold. \begin{enumerate} \item The above $\Phi$ is indeed a natural transformation from $\id_{\Modtb}$ to $\rr^\ast$. \item Let $\iota: \Modone \to \Modtb$ be the canonical embedding. Then $\Phi$ induces the natural isomorphism $\id_{\Modone} \Longrightarrow \rr^\ast \circ \iota$ between endofunctors of $\Modone$. \item The above $\Psi$ is indeed a natural transformation from $\ss^\ast$ to $\sigma_{\One - \tb}$. \end{enumerate} \end{Lemma} \begin{proof} We will prove only the assertions (1) and (2). The rest is proved by the way similar to (1). (1) First, we must verify that $\Phi_M$ is an $S$-linear map. Take any $u \in M_\ab$ and $\bb \in \ZZ^n$ with $\bb \ge \Zero$. Then \begin{align*} \phi^M_{\ab + \bb} (x^\bb \cdot \Phi_M(u)) & = x^{(\ab + \bb ) \cdot \tb - \rr(\ab + \bb)} (x^{\rr(\ab + \bb) - \rr(\ab)} {(\phi^M_{\ab})}^{-1}(x^{\ab \cdot (\tb - \One)}u)) \\ & = x^{(\ab + \bb) \cdot \tb - \rr(\ab)} {(\phi^M_{\ab})}^{-1}(x^{\ab \cdot (\tb - \One)}u) \\ & = x^{\bb \cdot \tb} (x^{\ab \cdot \tb - \rr(\ab)} {(\phi^M_{\ab})}^{-1}(x^{\ab \cdot (\tb - \One)}u)) = x^{\bb \cdot \tb} \cdot x^{\ab \cdot (\tb -\One)}u\\ & = x^{(\ab + \bb)\cdot (\tb - \One)} (x^\bb u), \end{align*} and hence it follows that $$ x^\bb \cdot \Phi_M(u) = (\phi^M_{\ab + \bb})^{-1}(x^{(\ab + \bb) \cdot (\tb - \One)} (x^\bb u)) = \Phi_M(x^\bb u). $$ Thus $\Phi_M$ is indeed $S$-linear. Next let $f: M \to N$ be a morphism in $\Modtb$. We will show that the following diagram commutes; $$ \begin{CD} M @>\Phi_M >> \rr^\ast(M) \\ @VfVV @VV\rr^\ast(f)V \\ N @>>\Phi_N> \rr^\ast(N). \end{CD} $$ Let $u \in M_\ab$. Then \begin{align*} \phi^N_\ab(\rr^\ast(f) \circ \Phi_M(u)) &= x^{\ab \cdot \tb - \rr(\ab)} \cdot f (\Phi_M(u)) \\ &= f(x^{\ab \cdot \tb - \rr(\ab)} \cdot \Phi_M(u) ) \\ &= f(x^{\ab \cdot \tb - \rr(\ab)} \cdot (\phi^M_\ab)^{-1}(x^{\ab \cdot (\tb - \One)} u)) \\ &= f(x^{\ab \cdot (\tb - \One)} u) = x^{\ab \cdot(\tb - \One)} \cdot f(u) \end{align*} Therefore we conclude that $$ \rr^\ast(f) \circ \Phi_M(u) = (\phi^N_\ab)^{-1}(x^{\ab \cdot (\tb - \One)} \cdot f(u)) = \Phi_N \circ f (u). $$ (2) Let $M \in \Modone$. We have to show $\Phi_M: M \to \rr^\ast(M)$ is then an isomorphism. Since both of $M$ and $\rr^\ast(M)$ are objects in $\Modone$, it suffices to show that each $(\Phi_M)_\ab: M_\ab \to \rr^\ast(M)_\ab$ is an isomorphism for all $\ab$ with $\Zero \le \ab \le \One$. This is an immediate consequence of the fact that for such $\ab$, the multiplication map $$ M_\ab \xrightarrow{x^{\ab \cdot (\tb - \One)}} M_{\ab \cdot \tb} $$ is an isomorphism since $M \in \Modone$. Thus we conclude that $\Phi_M$ is an isomorphism for all $M \in \Modone$. \end{proof} \begin{comment} \begin{Lemma}\label{help} For any $\ab$, $\bb\in \ZZ^n$, there is the following natural isomorphism of functors \[ \tau_{\ab}\circ \sigma_{\bb} \simeq \sigma_{\bb}\circ \tau_{\ab-\bb} \] from $\ModZ$ to itself. \end{Lemma} \begin{proof} Let $\cb\in \ZZ^n.$ Then \[ (\sigma_\bb\circ \tau_{\ab-\bb} (M))_{\cb}=(\tau_{\ab-\bb}(M)(-\bb))_\cb=\left\{ \begin{array}{ll} 0, & \cb-\bb\not\geq \ab-\bb,\\ M_{\cb-\bb},& \cb-\bb \geq \ab-\bb, \end{array} \right. \] By comparing with the degree $\cb$ component of $\tau_{\ab}\circ \sigma_{\bb}(M)$, the desired claim follows immediately. \end{proof} \end{comment} Let $\Dc$ denote the contravariant functor $\grHom_S(-, S): \ModZ \to \ModZ$. We set $\Dc_{\tb} := \sigma_{\tb } \circ \Dc$. Note that $\Dc_{\tb}$ gives a duality on $\Modtb$, and $\Dc_{\One} = \sigma_{-\tb + \One} \circ \Dc_{\tb}$ is the usual duality on $\ModZ$ by the canonical module $S(-\One)$ of $S$. The functor $\Dc_{\tb}$, lifted up to a functor from the category of complexes in $\Modtb$ to itself, coincides with the one $\Dc_\tb$ in \cite{BF} up to shifting and quasi-isomorphism \cite[Proposition 3.6]{BF}. Moreover $\Dc_\tb$ (resp. $\Dc_\One$) sends $M \in \Modtb$ (resp. $M \in \Modone$) to an object in $\Modtb$ (resp. $\Modone$). \begin{Proposition}\label{prop:r_and_D} There exists the natural isomorphisms between functors \begin{enumerate} \item $\tau_{\Zero} \circ \Dc_\One \simeq \Dc_\One \circ \rr^\ast$ and \item $\rr^\ast \circ \Dc_\tb \simeq \Dc_\One \circ \ss^\ast$, \end{enumerate} from $\Modtb$ to $\Modone$. \end{Proposition} \begin{proof} (1) By Lemma \ref{lem:nat}, there exists the natural transformation $\Phi:\id_{\Modtb} \Longrightarrow \rr^\ast$, and hence we have the one $\tau_{\Zero} \circ \Dc_{\One} \circ \rr^\ast \Longrightarrow \tau_{\Zero} \circ \Dc_{\One}$, where both functors are regarded as the ones from $\Modtb$ to $\ModN$. Since $\rr^\ast M$ is a squarefree module, it follows that $\tau_{\Zero} \circ \Dc_{\One} \circ \rr^\ast = \Dc_{\One} \circ \rr^\ast$. Consequently, the above natural transformation induces the one $\Phi': \Dc_{\One} \circ \rr^\ast \Longrightarrow \tau_{\Zero} \circ \Dc_{\One}$ of functors from $\Modtb$ to $\ModN$. Note that any $M \in \Modtb$ has a presentation $$ F_1 \longto F_0 \longto M \longto 0 $$ with $F_0, F_1$ free modules given by direct sums of finitely many copies of $S(-\ab)$ with $\Zero \le \ab \le \tb$. Since the functors $\rr^\ast, \tau_{\tb - \One}$ are exact and since $\Dc_\tb$ is left exact, we have the following commutative diagram with exact rows; $$ \begin{CD} 0 @>>> \Dc_\One \circ \rr^\ast(M) @>>> \Dc_\One \circ \rr^\ast(F_0) @>>> \Dc_\One \circ \rr^\ast(F_1) \\ @. @V{\Phi'}_M VV @V{\Psi'}_{F_0}VV @V{\Phi'}_{F_1}VV \\ 0 @>>> \tau_{\Zero} \circ \Dc_\One (M) @>>> \tau_{\Zero} \circ \Dc_\One(F_0) @>>> \tau_{\Zero} \circ \Dc_\One(F_1). \end{CD} $$ Thus what we have to show is that ${\Phi'}_{S(-\ab)}$ is an isomorphism for any $\ab \in \ZZ^n$ with $\Zero \le \ab \le \tb$. Let $N$ be the cokernel of ${\Phi'}_{S(-\ab)}: S(-\ab) \to S(-\sqrt{\ab}) = \rr^\ast(S(-\ab))$. The map ${\Phi'}_S(-\ab)$ is obviously injective, and hence the following sequence \[ 0\to S(-\ab)\to \rr^\ast(S(-\ab))=S(-\sqrt{\ab})\to N\to 0, \] is exact. By applying $\Dc_{\One}$, we obtain: \[ 0\to \Dc_{\One}(N) \to \Dc_{\One}\circ \rr^\ast(S(-\ab)) \to \Dc_{\One}(S(-\ab)) \] As obviously $\dim N\leq n-1,$ it follows that $\Dc_\One(N) = 0$. Therefore we get the following exact sequence \[\begin{CD} 0 @>>> \Dc_\One(S(-\sqrt{\ab})) @>{\Phi'}_{S(-\ab)}>> \tau_{\Zero} \circ \Dc_\One(S(-\ab)). \end{CD} \] Now $\Dc_\One(S(-\sqrt{\ab})) \cong S(-\One + \sqrt{\ab})$ and $\Dc_\One(S(-\ab)) \cong S(-\One + \ab)$. Easy observation shows that for $\bb \in \ZZ^n$ with $\bb \ge \Zero$, $S(-\One + \sqrt{\ab})_\bb \neq 0$ if and only if $\tau_{\Zero}(S(-\One + \ab))_\bb \neq 0$. Therefore it follows that ${\Phi'}_{S(-\ab)}$ is an isomorphism. (2) The assertion (2) can be proved in the way similar to (1). We have the natural transformation $$ \rr^\ast \circ \Dc_\tb = \rr^\ast \circ \Dc_\One \circ \sigma_{\One - \tb} \Longrightarrow \rr^\ast \circ \Dc_\One \circ \ss^\ast $$ between functors from $\Modtb$ to $\Modone$ by Lemma 2.1. Since $\Dc_\One \circ \ss^\ast$ sends $M \in \Modtb$ to an object in $\Modone$, there is a natural isomorphism $\rr^\ast \circ \Dc_\One \circ \ss^\ast \simeq \Dc_\One \circ \ss^\ast$ by Lemma 2.1 again. Consequently we have the natural transformation $\Psi':\rr^\ast \circ \Dc_\tb \Longrightarrow \Dc_\One \circ \ss^\ast$. By the argument similar as above, we have only to prove ${\Psi'}_{S(-\ab)}$ is an isomorphism for all $\ab$ with $\Zero \le \ab \le \tb$. By the definition of $\Psi$, $\Psi_{S(-\ab)}$ is injective. Applying $\rr^\ast \circ \Dc_\One$ to the exact sequence $$ 0 \longto \ss^\ast(S(-\ab)) \xrightarrow{\Psi_{S(-\ab)}} \sigma_{\One - \tb}(S(-\ab)) \longto M \longto 0, $$ where $M$ is the cokernel of $\Psi_{S(-\ab)}$, we have the exact one $$ 0 \longto \rr^\ast \circ \Dc_\tb(S(-\ab)) \xrightarrow{{\Psi'}_{S(-\ab)}}\Dc_\One \circ \ss^\ast(S(-\ab)) $$ We define $\bb = (b_i)_{1 \le i \le n}$ by setting $b_i = 0$ if $a_i \le t_i - 1$ and $b_i = 1$ if $a_i = t_i$. It follows that $\ss^\ast(S(-\ab)) \cong S(-\bb)$, and hence $\Dc_\One \circ \ss^\ast(S(-\ab)) \cong S(-(\One - \bb))$. On the other hand, $\rr^\ast \circ \Dc_\tb(S(-\ab)) \cong S(-\sqrt{\tb - \ab})$. Easy calculation shows $\One - \bb = \sqrt{\tb - \ab}$, and therefore ${\Psi'}_{S(-\ab)}$ is an isomorphism. \end{proof} \begin{Theorem} \label{ext} The following statements hold. \begin{enumerate} \item Let $\omega_S$ be the canonical module of $S$. Then there are the following two isomorphisms \begin{enumerate} \item $\tau_{\Zero} \grExt^p_S(M, \omega_S) \cong \grExt^p_S(\rr^\ast(M), \omega_S)$ \item $\rr^\ast(\grExt^p_S(M, S(-\tb))) \cong \grExt^p_S(\ss^\ast(M),\omega_S)$ \end{enumerate} for all $p$ and $M \in \Modtb$. \item Assume $I$ is a Cohen-Macaulay monomial ideal such that $S/I \in \Modtb$. Let $\omega_{S/I}$, $\omega_{S/\sqrt I}$ be the canonical modules of $S/I$, $S/\sqrt I$, respectively. Then $$ \omega_{S/\sqrt I} \cong \tau_{ \Zero}(\omega_{S/I}) = (\omega_{S/I})_{\ge \Zero}. $$ \end{enumerate} \end{Theorem} \begin{proof} Choose a $\ZZ^n$-graded minimal free resolution $P_\bullet$ of $M$ with each $P_i$ positively $\tb$-determined. By Proposition~\ref{prop:r_and_D}, $$ \tau_\Zero \grExt^p_S(M,\omega_S) \cong H^p(\tau_\Zero \circ \Dc_\One(P_\bullet)) \cong H^p(\Dc_\One \circ \rr^\ast(P_\bullet)) \cong \grExt^p(\rr^\ast(M), \omega_S), $$ since $\rr^\ast(P_\bullet)$ is a free resolution of $\rr^\ast(M)$. The assertion (2) follows from (1) and Theorem~\ref{CM}. \end{proof} \begin{Remark}{\em In \cite[Corollary 2.3]{HTT}, Herzog, Takayama, and Terai proved $$ H^i_\mm(S/I)_\ab \cong H^i_\mm(S/\sqrt{I})_\ab $$ for any monomial ideal $I$ and all $i$ and $\ab \le \Zero$. In (1) of Theorem \ref{ext}, taking the graded Matlis duality, we obtain the generalization of this result, which also implies that there exists the isomorphism $$ H^i_\mm(S/I)_{\le \Zero} \cong H^i_\mm(S/\sqrt I) $$ as $\ZZ^n$-graded $S$-modules. Here, for any $M \in \ModZ$, $M_{\le \Zero}$ denotes the residue of $M$ by its $\ZZ^n$-graded submodule generated by the homogeneous elements whose degree is not less than or equal to $\Zero$.} \end{Remark} As an immediate consequence of the above theorem we get the following corollary that generalizes a result of \cite{HTT}. \begin{Corollary}\label{gCM} Let $M$ be a positively $\tb$-determined $S$-module with $\dim M=\dim \rr^\ast M.$ Then: \begin{enumerate} \item $M$ is generalized Cohen-Macaulay if and only if so is $\rr^\ast M$. \item If $M$ is Buchsbaum, then $\rr^\ast M$ is Buchsbaum. \end{enumerate} \end{Corollary} \begin{proof} (1) $M$ is generalized Cohen-Macaulay if and only if $\grExt^i_S(M,\omega_S)$ has finite length for any $i\neq n-d$, where $d=\dim M$. This is equivalent to say that $\tau_0 \grExt^i_S(M,\omega_S)$ has finite length for $i\neq n-d$. Therefore, since $M$ and $\rr^\ast M$ have the same dimension, the desired statement follows by Corollary~\ref{ext}. (2) If $M$ is Buchsbaum, then $M$ is generalized Cohen-Macaulay \cite{Go}, thus $\rr^\ast M$ is generalized Cohen-Macaulay. By \cite[Corollary 2.7]{Y}, it follows that $\rr^\ast M$ is Buchsbaum. \end{proof} \section{The $\rr^\ast$ functor and Alexander duality} Recall that there exists the duality $\Ac_{\tb}$ on $\Modtb$, called Alexander duality, defined by Miller \cite{Mill}. In the case, $\tb = \One$, R\"{o}mer also defines independently in \cite{R}. Let $E$ denote the injective hull of $K$ and $p_\tb: \NN^n \to \ZZ^n$ the map defined by $p_\tb(\ab) := (p_{t_1}(a_1),\dots ,p_{t_n}(a_n))$, where $$ p_{t_i}(a_i) := \begin{cases} a_i & \text{if $0 \le a_i < t_i$,} \\ t_i & \text{if $a_i \ge t_i$.} \end{cases} $$ The functor $\Ac_{\tb}$ is given by the formula $$ \Ac_{\tb}(M) = p_{\tb}^\ast\grHom_S(M(\tb),E). $$ If we set $\DD$ to be the functor $\grHom_K(-, K)$, we have the natural isomorphism $\Ac_\tb \cong p_\tb^\ast \circ \sigma_\tb \circ \DD$. The following is essentially proved by Miller in \cite{Mill}. \begin{Proposition}\label{prop:r_and_A} There exists a natural isomorphism of functors from $\Modtb$ to itself $$ \Ac_\One \circ \rr^\ast \simeq \rr^\ast \circ \Ac_{\tb}. $$ \end{Proposition} \begin{proof} Let $M \in \Modtb$ and $\ab \in \ZZ^n$ with $\ab \ge \Zero$. \begin{align*} \Ac_\One \circ \rr^\ast (M)_\ab &= \grHom_K( \rr^\ast(M), K)_{p_\One(\ab) - \One} = \grHom_K(M_{\rr(- p_\One(\ab) + \One)}, K), \\ \rr^\ast \circ \Ac_\tb (M)_\ab &= \grHom_K(M, K)_{p_\tb(\rr(\ab)) - \tb} = \grHom_K(M_{-p_\tb(\rr(\ab)) + \tb}, K). \end{align*} It is easy to verify that $\rr(- p_\One(\ab) + \One) = -p_\tb(\rr(\ab)) + \tb$. Hence $\Ac_\One \circ \rr^\ast(M) \cong \rr^\ast \circ \Ac_\tb(M)$ as $\ZZ^n$-graded $K$-vector spaces. By a routine calculation, we can show that this isomorphism is $S$-linear and natural. \end{proof} \section{Relation to the Auslander-Reiten translate} Let $\le$ be the order on $\ZZ^n$ defined as follows: $$ \ab \le \bb \ \Longleftrightarrow \ a_i \le b_i $$ for all $i$. With the order induced by $<$, the set $P_\tb := \{ \ab \in \ZZ^n : \Zero \le \ab \le \tb \}$ becomes a poset. Let $A$ denote the incidence algebra of $P_\tb$ over $K$. It is well-known that the category $\Modtb$ is equivalent to the one $\mod A$ consisting of finite-dimensional left $A$-modules (\cite[3.5]{BF} and \cite[Proposition 4.3]{Y2}). Since $A$ is a finite-dimensional $A$-algebra of finite global dimension, as is shown in \cite[3.6]{H} by Happel, the bounded derived category $D^b(\mod A)$ of $\mod A$ has Auslander-Reiten triangles. Let $K^b(\Proj A)$ (resp. $K^b(\Inj A)$) be the bounded homotopy category of complexes of finite-dimensional projective (resp. injective) left modules. According to Happel's proof, through the equivalence $K^b(\Proj A) \cong D^b(\mod A) \cong K^b(\Inj A)$ of triangulated categories, the Auslander-Reiten translate (see \cite[Definition 1.2]{J} for its definition) is then given by $T^{-1} \circ v$, where $T$ is the usual translation functor and $v$ is the equivalence $K^b(\Proj A) \cong K^b(\Inj A)$ of triangulated categories induced by the Nakayama functor, i.e., $\Hom_K(\Hom_A(-,A),K)$. On the other hand $v$ is coincides with $\Ac_\tb \circ \Dc_\tb$ through the equivalence $\Modtb \cong \mod A$. In this sense, $\Ac_\tb \circ \Dc_\tb$ represents the Auslander-Reiten translate of $D^b(\mod A)$ (see\cite[3.3 and 3.5]{BF} for details). In this section, we discuss the relation between $\rr^\ast$ and $\Ac_\tb \circ \Dc_\tb$. For this, we need to define a new functor. For $\ab \in \ZZ^n$, let $\tau^\ab$ be the functor from $\ModZ$ to $\ModZ$ defined as follows: for any $M \in \ModZ$, we set $$ \tau_\ab(M) := M/(S \cdot \bigoplus_{\bb \not\le \ab} M_\ab), $$ and for a morphism $f: M \to N$ in $\ModZ$, we assign, $\tau_\ab(f)$, the natural homomorphism induced by $f$. Clearly $\tau_\ab$ is additive and exact. \begin{Lemma}\label{lem:sig_tau_2} There are the following natural isomorphisms of functors: for any $\ab$, $\bb \in \ZZ^n$, \begin{enumerate} \item $\sigma_\ab \circ \tau^\bb \simeq \tau^{\ab + \bb} \circ \sigma_\bb$, \item $\DD \circ \sigma_\ab \simeq \sigma_{-\ab} \circ \DD$, and \item $\DD \circ \tau_\ab \simeq \tau^\ab \circ \DD$. \end{enumerate} \end{Lemma} \begin{proof} By comparing each degree $\cb$ component, for each $M \in \ModZ$, we obtain an isomorphism, as $\ZZ^n$-graded $K$-vector spaces, between two modules given by applying each of two functors. An easy calculation shows these maps indeed defines the desired natural isomorphisms. \end{proof} \begin{Proposition}\label{ART} There exists the following two natural isomorphisms of functors \begin{enumerate} \item $\Ac_\One \circ \Dc_\One \circ \rr^\ast \simeq p^\ast_\One \circ \Ac_\tb \circ \Dc_\tb$ \item $\rr^\ast \circ \Ac_\tb \circ \Dc_\tb \simeq \Ac_\One \circ \Dc_\One \circ \ss^\ast$ \end{enumerate} from $\Modtb$ to $\Modone$. \end{Proposition} \begin{proof} The second natural isomorphism is a direct consequence of Propositions \ref{prop:r_and_D} and \ref{prop:r_and_A}. We will show the first. It follows from Proposition~\ref{prop:r_and_D} and Lemma~\ref{lem:sig_tau_2} that \begin{align*} \Ac_\One \circ \Dc_\One \circ \rr^\ast &\simeq \Ac_\One \circ \tau_\Zero \circ \Dc_\One \\ &\simeq p^\ast_\One \circ \sigma_\One \circ \DD \circ \tau_\Zero \circ \sigma_{\One - \tb} \circ \Dc_\tb \\ &\simeq p^\ast_\One \circ \sigma_{\One} \circ \tau^{\Zero} \circ \sigma_{\tb - \One} \circ \DD \circ \Dc_\tb \\ &\simeq p^\ast_\One \circ \tau^\One \circ \sigma_\tb \circ \DD \circ \Dc_\tb. \end{align*} By the definition, it is easy to verify that $p^\ast_\One \circ \tau^\One = p^\ast_\One$. Moreover $p_\tb \circ p_\One = p_\One$ implies $p^\ast_\One \circ p^\ast_\tb = p^\ast_\One$. Thus it follows that $$ \Ac_\One \circ \Dc_\One \circ \rr^\ast \simeq p^\ast_\One \circ \tau^\One \circ \sigma_\tb \circ \DD \circ \Dc_\tb \simeq p^\ast_\One \circ \Ac_\tb \circ \Dc_\tb. $$ \end{proof}
\section{Introduction} A \emph{\mbox{\ensuremath{\beta(1,0)}-tree}}~\cite{JaGi98} is a rooted plane tree labeled with positive integers such that \begin{enumerate} \item Leaves have label $1$. \item The root has label equal to the sum of its children's labels. \item Any other node has label no greater than the sum of its children's labels. \end{enumerate} Below is an example of such a tree.\\[-3ex] \extree In~\cite{CKS} we introduced an involution, $h$, on {\btree s}. We also gave a result on the equidistribution of certain statistics on {\btree s}. A proof that $h$ indeed is an involution was, however, not given; rather, the proof was said to be found in a forthcoming paper that never materialised. The proof of the equidistribution was in fact also omitted. In this note we give the two missing proofs. We also refine the equidistribution result. \section{The structure of {\btree s}} We say a {\mbox{\ensuremath{\beta(1,0)}-tree}} on two or more nodes is \emph{indecomposable} if its root has exactly one child and \emph{decomposable} if it has more than one child. The {\mbox{\ensuremath{\beta(1,0)}-tree}} on one node, $\leaf$ = $\bx\!$, is neither indecomposable nor decomposable. Let $\B_n$ be the set of all {\btree s} on $n$ nodes, and let $\barB_n$ be the subset of $\B_n$ consisting of the indecomposable trees. Let $\B_n^k$ be the subset of $\B_n$ consisting of the trees with root label $k$. For instance, $$ \B_3=\big\{\;\,\bbbx\;,\;\bbbxx\;\big\}\qquad\; \barB_3=\B_3^1=\big\{\;\bbbx\;\;\,\big\}\qquad\; \B_3^2=\big\{\;\bbbxx\;\big\} $$ Decomposable trees can be regarded as sums of indecomposable ones: \begin{center} \begin{tikzpicture}[ scale=0.45, baseline=0pt ] \style \exampletree \end{tikzpicture} $\quad = \quad$ \begin{tikzpicture}[ scale=0.45, baseline=0pt ] \style \path node [disc] (r1) at (-2, 0) {} node [disc] (r2) at ( 2, 0) {}; \exampleforest \draw (r1) node[above left=-1pt] {1} -- (1) (r2) node[above right=-1pt] {3} -- (2); \node[font=\normalsize] at (0,0) {$\oplus$}; \end{tikzpicture} \end{center} \smallskip In fact we do not need to require $u$ and $v$ to be indecomposable for the sum $u\oplus v$ to make sense. In general, we define that the root label of $u\oplus v$ is the sum of the root label of $u$ and the root label of $v$, and that the subtrees of $u\oplus v$ are those of $u$ followed by those of $v$. So, $$ \begin{tikzpicture}[ scale=0.45, baseline=(r.base) ] \style \node [disc] (r) at (0,1) {}; \node [disc] (1) at (-1,0) {}; \draw (r) node[above=1pt] {\ensuremath{1}} -- (1) node[below=1pt] {\ensuremath{1}}; \end{tikzpicture}\oplus \begin{tikzpicture}[ scale=0.45, baseline=(r1.base) ] \style \node [disc] (r1) at ( 0,-1) {}; \node [disc] (r11) at ( 1,-2) {}; \node [disc] (r12) at ( 0,-2) {}; \draw (r1) -- (r11) node[below=1pt] {1} (r1) -- (r12) node[below=1pt] {1}; \draw (r1) node[above=1pt] {2}; \end{tikzpicture} = \begin{tikzpicture}[ scale=0.45, baseline=(r1.base) ] \bxxx \draw (r1) node[above=1pt] {3}; \end{tikzpicture} = \begin{tikzpicture}[ scale=0.45, baseline=(r1.base) ] \style \node [disc] (r1) at ( 0,-1) {}; \node [disc] (r11) at (-1,-2) {}; \node [disc] (r12) at ( 0,-2) {}; \draw (r1) -- (r11) node[below=1pt] {1} (r1) -- (r12) node[below=1pt] {1}; \draw (r1) node[above=1pt] {2}; \end{tikzpicture} \oplus \begin{tikzpicture}[ scale=0.45, baseline=(r.base) ] \style \node [disc] (r) at (0,1) {}; \node [disc] (1) at (1,0) {}; \draw (r) node[above=1pt] {\ensuremath{1}} -- (1) node[below=1pt] {\ensuremath{1}}; \end{tikzpicture} $$ Further, there is a simple one-to-one correspondence $\lambda$ between the Cartesian product $[k]\times\B_{n-1}^k$ and the disjoint union $\cup_{i=1}^k\barB_{n}^i$, where $\barB_n^k$ is the subset of $\barB_n$ consisting of the trees with root label $k$: $$ \begin{tikzpicture}[ scale=0.45, baseline=(r11.base) ] \bxxx \draw (r1) node[above=1pt] {3}; \end{tikzpicture} \,\raisebox{2ex}{$\substack{\lambda_1\\ \longrightarrow}$}\, \begin{tikzpicture}[ scale=0.45, baseline=(r11.base) ] \bxxx \node [disc] (r) at (0,0) {}; \draw (r) node[above right=-1pt] {1} -- (r1) node[above right=-1pt] {1}; \end{tikzpicture} \qquad\;\; \begin{tikzpicture}[ scale=0.45, baseline=(r11.base) ] \bxxx \draw (r1) node[above=1pt] {3}; \end{tikzpicture} \,\raisebox{2ex}{$\substack{\lambda_2\\ \longrightarrow}$}\, \begin{tikzpicture}[ scale=0.45, baseline=(r11.base) ] \bxxx \node [disc] (r) at (0,0) {}; \draw (r) node[above right=-1pt] {2} -- (r1) node[above right=-1pt] {2}; \end{tikzpicture} \qquad\;\; \begin{tikzpicture}[ scale=0.45, baseline=(r11.base) ] \bxxx \draw (r1) node[above=1pt] {3}; \end{tikzpicture} \,\raisebox{2ex}{$\substack{\lambda_3\\ \longrightarrow}$}\, \begin{tikzpicture}[ scale=0.45, baseline=(r11.base) ] \bxxx \node [disc] (r) at (0,0) {}; \draw (r) node[above right=-1pt] {3} -- (r1) node[above right=-1pt] {3}; \end{tikzpicture} $$ In general, if $t$ is a tree with root label $k$ and $i$ is an integer such that $1\leq i\leq k$, then $\lambda_i t$ is obtained from $t$ by joining a new root via an edge to the old root; and both the new root and the old root are assigned the label $i$. Thus each \mbox{\ensuremath{\beta(1,0)}-tree}, $t$, is of exactly one the following three forms: \begin{itemize} \item[] $t=\leaf$, \hfill (the single node tree) \item[] $t=u\oplus v$, \hfill (decomposable) \item[] $t=\lambda_i u$, where $1\leq i\leq \root u$,\hfill(indecomposable) \end{itemize} in which $u$ and $v$ are \btree s, and $\root u$ denotes the root label of $u$. As an example of the encoding this characterisation entails we have $$ \begin{tikzpicture}[ scale=0.45, baseline=(111.base) ] \style \path node [disc] (1) at ( 0, -1) {} node [disc] (11) at ( 0, -2) {} node [disc] (111) at (-.65, -3) {} node [disc] (112) at (0.65, -3) {}; \draw (1) node[right=2pt] {2} -- (11) node[right=2pt] {2} (11) -- (111) node[below=1pt] {1} (11) -- (112) node[below=1pt] {1}; \end{tikzpicture} = \lambda_2\Big( \begin{tikzpicture}[ scale=0.45, baseline=(111.base) ] \style \path node [disc] (11) at ( 0, -2) {} node [disc] (111) at (-0.65, -3) {} node [disc] (112) at ( 0.65, -3) {}; \draw (11) node[above=2pt] {2} -- (111) node[below=1pt] {1} (11) -- (112) node[below=1pt] {1}; \end{tikzpicture} \Big) = \lambda_2\Big(\bbx\oplus\bbx\,\Big) = \lambda_2\Big(\lambda_1(\leaf)\oplus\lambda_1(\leaf)\,\Big) $$ \section{An involution on {\btree s}}\label{h} In this section we define an involution on {\btree s}. To that end we now describe a new way of decomposing {\btree s}. Schematically the sum $\oplus$ on {\btree s} is described by \begin{center} \makebox[3.3cm][l]{ \begin{tikzpicture}[semithick, scale=0.3] \def\tri{ -- +(1,-1.732) -- +(-1,-1.732) -- cycle}; \draw (0,0) \tri; \filldraw[fill=white] (0,0) circle (2mm); \node[font=\footnotesize] at (0,0.7) {$a$}; \node at (1.5,-0.6) {$\oplus$}; \filldraw (3,0) \tri; \filldraw[fill=white] (3,0) circle (2mm); \node[font=\footnotesize] at (3,0.8) {$b$}; \node at (5,-0.6) {=}; \draw[shift={(8,0)}] (0,0) [rotate=-45] \tri; \filldraw[shift={(8,0)}] (0,0) [rotate=45 ] \tri; \filldraw[fill=white] (8,0) circle (2mm); \node[font=\footnotesize] at (8,0.8) {$a+b$}; \end{tikzpicture} } \end{center} An alternative sum is\\[-5ex] \begin{center} \makebox[3.3cm][l]{ \begin{tikzpicture}[semithick, scale=0.3] \def\tri{ -- +(1,-1.732) -- +(-1,-1.732) -- cycle}; \draw (0,0) \tri; \filldraw[fill=white] (0,0) circle (2mm); \node[font=\footnotesize] at (0,0.7) {$a$}; \node at (1.5,-0.6) {$\obslash$}; \filldraw (3,0) \tri; \filldraw[fill=white] (3,0) circle (2mm); \node[font=\footnotesize] at (3,0.8) {$b$}; \node at (5,-0.6) {=}; \draw (7,0) \tri; \filldraw (8,-1.732) \tri; \filldraw[fill=white] (7,0) circle (2mm); \filldraw[fill=white] (8,-1.732) circle (2mm); \node[font=\footnotesize] at (7,0.7) {$a$}; \node[font=\footnotesize] at (8.7,-1.732) {$1$}; \end{tikzpicture} } \end{center} That is, to get $u\obslash v$ we join $u$ and $v$ by identifying the rightmost leaf in $u$ with the root of $v$, and that node is assigned the label $1$. The \emph{right path} is the path from the root to the rightmost leaf. Let $\rpath(t)$ denote the length of (number of edges on) the right path of $t$. Note that \begin{align} \root(u\oplus v) &= \root u + \root v\\ \rpath(u\oplus v) &= \rpath v \\ \shortintertext{while} \root(u\obslash v) &= \root u \\ \rpath(u\obslash v) &= \rpath u + \rpath v \label{rpath_obslash}. \end{align} for $u\neq\leaf$ and $v\neq\leaf$. Thus, with respect to $\obslash$, $\rpath$ plays the role of $\root$, and vice versa. There is also a map $\gamma$ that plays a role analogous to that of $\lambda$: $$ \begin{tikzpicture}[scale=0.45, baseline=(r11.base) ] \eeev; \draw (r) node[above left=-0.5pt] {1} -- (r1) node[above left=-0.5pt] {1} -- (r11) node[above left=-0.5pt] {2}; \end{tikzpicture} \;\;\raisebox{2ex}{$\substack{\gamma_1\\ \longrightarrow}$}\; \begin{tikzpicture}[ scale=0.45, baseline=(r11.base) ] \eeev; \node [disc] (r2) at ( 0.8, 2 ) {}; \draw (r) node[above right=-0.5pt] {2} -- (r1) node[above left=-0.5pt] {1} -- (r11) node[above left=-0.5pt] {2}; \draw (r) -- (r2) node[below=1pt] {1}; \end{tikzpicture} \qquad\quad\; \begin{tikzpicture}[scale=0.45, baseline=(r11.base) ] \eeev; \draw (r) node[above left=-0.5pt] {1} -- (r1) node[above left=-0.5pt] {1} -- (r11) node[above left=-0.5pt] {2}; \end{tikzpicture} \;\;\raisebox{2ex}{$\substack{\gamma_2\\ \longrightarrow}$}\; \begin{tikzpicture}[ scale=0.45, baseline=(r11.base) ] \eeev; \node [disc] (r12) at ( 0.8, 1 ) {}; \draw (r) node[above right=-0.5pt] {2} -- (r1) node[above right=-0.5pt] {2} -- (r11) node[above left=-0.5pt] {2}; \draw (r1) -- (r12) node[below=1pt] {1}; \end{tikzpicture} \qquad\quad\; \begin{tikzpicture}[scale=0.45, baseline=(r11.base) ] \eeev; \draw (r) node[above left=-0.5pt] {1} -- (r1) node[above left=-0.5pt] {1} -- (r11) node[above left=-0.5pt] {2}; \end{tikzpicture} \;\;\raisebox{2ex}{$\substack{\gamma_3\\ \longrightarrow}$}\; \begin{tikzpicture}[ scale=0.45, baseline=(r11.base) ] \eeev; \node [disc] (r112) at ( 0.8, 0.1 ) {}; \draw (r) node[above right=-0.5pt] {2} -- (r1) node[above right=-0.5pt] {2} -- (r11) node[above right=-0.5pt] {3}; \draw (r11) -- (r112) node[below=1pt] {1}; \end{tikzpicture} $$ Here is how $\gamma_i t$ is defined in general: Assume that the length of the right path of $t$ is $k$ and that $i$ is an integer such that $1\leq i\leq k$. Let us by $x$ refer to the $i$th node on the right path of $t$. Then $\gamma_i t$ is obtained from $t$ by joining a new leaf via an edge to $x$, making the new leaf the rightmost leaf in $\gamma_i t$; and, lastly, add $1$ to the label of each node on the right path, except for the new leaf. Note that $\rpath\gamma_i t = i$. We explore the two ways to decompose {\btree s} we now have by defining an endofunction $h:\B\to\B$ as follows: \begin{align*} h(\leaf) &= \leaf; \\ h(\lambda_it) &= \gamma_ih(t); \\ h(u\oplus v) &= h(v) \obslash h(u). \end{align*} For instance, $$ \begin{tikzpicture}[ scale=0.35, baseline=-7.5ex ] \discstyle \exampletree \end{tikzpicture}\, \begin{aligned} \;=\;\;& \lambda_1\Big(\lambda_2\big(\,\edge\oplus\edge\,\big)\oplus\edge\,\Big)\oplus \lambda_3\big(\,\edge\oplus\edge\oplus\edge\,\big)\\ \raisebox{.6ex}{$\substack{\textstyle{h}\\ \textstyle{\to}}$}\;& \gamma_3\big(\,\edge\obslash\edge\obslash\edge\,\big) \obslash\gamma_1\Big(\,\edge\obslash\gamma_2\big(\,\edge\obslash\edge\,\big)\Big) \;=\, \end{aligned} \begin{tikzpicture}[scale=0.35, baseline=(r1.base) ] \style \node [disc] (r) at ( 0, 7.4 ) {}; \node [disc] (r1) at ( 0, 6.2 ) {}; \node [disc] (r11) at ( 0, 5 ) {}; \node [disc] (r111) at (-0.8, 4 ) {}; \node [disc] (r112) at ( 0.8, 4 ) {}; \node [disc] (r1121) at ( 0, 3 ) {}; \node [disc] (r11211) at ( 0, 1.8 ) {}; \node [disc] (r112111) at (-0.8, 0.8 ) {}; \node [disc] (r112112) at ( 0.8, 0.8 ) {}; \node [disc] (r1122) at ( 1.6, 3 ) {}; \draw (r) node[above=1pt] {2} -- (r1) node[above right=-1pt] {2} (r1) -- (r11) node[above right=-.8pt] {2} (r11) -- (r111) node[left=1pt] {1} (r11) -- (r112) node[above right=-.8pt] {1} (r112) -- (r1121) node[left=1pt] {1} (r1121) -- (r11211) node[above right=-.8pt] {2} (r11211) -- (r112111) node[below=1pt] {1} (r11211) -- (r112112) node[below=1pt] {1} (r112) -- (r1122) node[below=1pt] {1}; \end{tikzpicture} $$ We will soon see that $h$ is in fact an involution! First we state some almost self-evident lemmas about relations between $\oplus$, $\obslash$, $\lambda$, and $\gamma$. \begin{lemma}\label{LG} Let $t$, $u$, and $v$ be {\btree s}. Then $$t\oplus(u \obslash v) = (t\oplus u) \obslash v. $$ \end{lemma} \begin{lemma}\label{lG&gL} Let $u$ and $v$ be {\btree s}. Then \begin{align*} \lambda_i(u \obslash v) &= (\lambda_iu)\obslash v\/;\\ \gamma_i(u\oplus v) &= u\oplus(\gamma_iv). \end{align*} \end{lemma} \begin{lemma}\label{lg} Let $t$ be a {\mbox{\ensuremath{\beta(1,0)}-tree}}. Then \begin{align*} \gamma_1 t &= t \oplus \edge\,; \\ \lambda_1 t &= \edge \obslash t\,; \\ \gamma_{i+1}\lambda_{j} &= \lambda_{j+1}\gamma_{i}. \end{align*} \end{lemma} Next we apply the lemmas above to prove the following lemma which is the most crucial component in establishing that $h$ is an involution. \begin{lemma}\label{h-dual} Let $t$, $u$, and $v$ be {\btree s}. Then $$h(\leaf) = \leaf,\;\; h(\gamma_it) = \lambda_ih(t),\,\text{ and }\; h(u\obslash v) = h(v)\oplus h(u). $$ \end{lemma} \begin{proof} We use induction on the number of nodes. The base case is trivial. The proof of the second claim is split into two cases:\\ Case 1, $t=\lambda_ju$: We shall prove that $h(\gamma_i\lambda_ju) = \lambda_i h(\lambda_ju)$ for all positive integers $i$ and $j$. If $i=1$, then\smallskip \begin{align*} h(\gamma_1\lambda_j u) &= h(\lambda_j u\oplus \edge) &&\text{by Lemma \ref{lg}}\\ &= h(\,\edge\,)\obslash h(\lambda_j u) &&\text{by definition of $h$}\\ &= \edge \obslash \gamma_jh(u) &&\text{by definition of $h$}\\ &= \lambda_1\gamma_j h(u) &&\text{by Lemma \ref{lg}}\\ &= \lambda_1h(\lambda_j u) &&\text{by definition of $h$} \shortintertext{If $i>1$, then} h(\gamma_i\lambda_ju) &= h(\lambda_{j+1}\gamma_{i-1}u) &&\text{by Lemma \ref{lg}}\\ &= \gamma_{j+1}h(\gamma_{i-1}u) &&\text{by definition of $h$}\\ &= \gamma_{j+1}\lambda_{i-1}h(u) &&\text{by induction}\\ &= \lambda_i\gamma_jh(u) &&\text{by Lemma \ref{lg}}\\ &= \lambda_i h(\lambda_ju) &&\text{by definition of $h$}\\ \shortintertext{Case 2, $t=u\oplus v$:} h\gamma_i(u\oplus v) &= h(u\oplus\gamma_iv) &&\text{by Lemma \ref{lG&gL}}\\ &= h(\gamma_iv) \obslash h(u) &&\text{by definition of $h$}\\ &= \lambda_i h(v) \obslash h(u) &&\text{by induction}\\ &= \lambda_i\big(h(v) \obslash h(u)\big) &&\text{by Lemma \ref{lG&gL}}\\ &= h(u\oplus v) &&\text{by definition of $h$} \end{align*} The proof of the third claim is also split into two cases.\\ Case 1, $u=\lambda_it$:\smallskip \begin{align*} h(\lambda_it \obslash v) &= h\lambda_i(t \obslash v) &&\text{by Lemma \ref{lG&gL}}\\ &= \gamma_ih(t \obslash v) &&\text{by Lemma \ref{h-dual}}\\ &= \gamma_i\big(h(t)\oplus h(v)\big) &&\text{by induction}\\ &= h(v)\oplus\gamma_ih(t) &&\text{by Lemma \ref{lG&gL}}\\ &= h(v)\oplus h(\lambda_it) &&\text{by definition of $h$} \shortintertext{Case 2, $u=s\oplus t$:} h((s \oplus t) \obslash v) &= h (s\oplus (t \obslash v)) &&\text{by Lemma \ref{LG}}\\ &= h(t \obslash v) \obslash h(s) &&\text{by definition of $h$}\\ &= \big(h(v)\oplus h(t)\big) \obslash h(s) &&\text{by induction}\\ &= h(v)\oplus\big(h(t) \obslash h(s)\big) &&\text{by Lemma \ref{LG}}\\ &= h(v)\oplus h\big(s\oplus t \big) &&\text{by definition of $h$} \end{align*} \end{proof} \begin{theorem} The function $h$ is an involution. \end{theorem} \begin{proof} We proceed by induction. By definition $h(\leaf)=\leaf$; using that twice the base case follows. For the induction step we consider indecomposable and decomposable trees separately. First, for indecomposable trees: $$ h^2(\lambda_it) = h\big(\gamma_ih(t)\big) = \lambda_i h^2(t)=\lambda_i(t). $$ Here we have used the definition of $h$, Lemma~\ref{h-dual}, and the induction hypothesis. Second, for decomposable trees: $$ h^2(u\oplus v) = h\big(h(v)\obslash h(u)\big) = h^2(v)\oplus h^2(u) = v\oplus u. $$ Again, we used the definition of $h$, Lemma~\ref{h-dual}, and the induction hypothesis. \end{proof} \section{Statistics on {\btree s}}\label{tree_stats} Let $t$ be a \mbox{\ensuremath{\beta(1,0)}-tree}. Recall that by $\root t$ we denote the label of the root. Recall also that the \emph{right path} is the path from the root to the rightmost leaf, and that the length of the right path is denoted $\rpath t$. By $\leaves t$ we denote the number of leaves in $t$; by $\internal t$ we denote the number of internal nodes (or nonleaves) in $t$. Note that the root is an internal node. The number of subtrees (or, equivalently, the number of children of the root) is denoted $\sub t$. Further, the number of $1$'s below the root on the right path is denoted $\rsub t$. \begin{theorem}\label{thm_h} On {\btree s} with at least one edge, the involution $h$ sends the first tuple below to the second. $$ \begin{array}{lllllllll} ( &\leaves, &\internal, &\root, &\rpath, &\sub, &\rsub & ) \\ ( & \internal, &\leaves, &\rpath, &\root, &\rsub, &\sub & ) \end{array} $$ \end{theorem} \begin{proof} We shall use induction to show that $\rpath h(t)=\root t$ and that $\root h(t)=\rpath t$; the other claims follow similarly. The base case is trivial. Assume that $t=\lambda_iu$ is indecomposable. Then $$ \rpath h(\lambda_iu) = \rpath\gamma_ih(u) = i = \root\lambda_iu $$ by definition of $h$, definition of $\root$ and $\rpath$, respectively. Furthermore, for a decomposable tree $t=u\oplus v$ we have \begin{align*} \rpath h(u\oplus v) &= \rpath \big(h(u) \obslash h(v)\big) && \text{by definition of $h$} \\ &= \rpath h(u) + \rpath h(v) && \text{by \eqref{rpath_obslash}} \\ &= \root u + \root v && \text{by induction} \\ &= \root (u\oplus v) && \text{by definition of $\root$} \end{align*} We have thus shown that $\rpath h(t) = \root t$ for any {\mbox{\ensuremath{\beta(1,0)}-tree}} $t$. That $\root h(t) = \rpath t$ follows from $h$ being an involution. \end{proof} The above theorem can be strengthened by introducing what we call labeled \btree s. $$ \begin{tikzpicture}[ scale=0.45, baseline=(111.base) ] \style \path node [disc] (1) at ( 0, -1) {} node [disc] (11) at ( 0, -2) {} node [disc] (111) at (-.65, -3) {} node [disc] (112) at (0.65, -3) {}; \draw (1) node[right=2pt] {$(2,a)$} -- (11) node[right=2pt] {$(2,b)$} (11) -- (111) node[below=1pt] {$(1,c)$\;\;} (11) -- (112) node[below=1pt] {\;\;$(1,d)$}; \end{tikzpicture} $$ This is a \mbox{\ensuremath{\beta(1,0)}-tree}\ in which each node has been assigned a unique label. In this example, the labels are taken from the alphabet $\{a,b,c,d\}$. A recursive characterization of labeled {\btree s} reads as follows. A \emph{labeled \mbox{\ensuremath{\beta(1,0)}-tree}} is of exactly one of the three forms: \begin{enumerate} \setcounter{enumi}{-1} \item $(1,x)$, a leaf with label $x$; \item $\lambda((i,x),t)$, where $t$ is a labeled {\mbox{\ensuremath{\beta(1,0)}-tree}} and $i\leq\root t$; \item $u \oplus v$, where $u$ and $v$ are labeled {\btree s}. \end{enumerate} Here we assume that the function $\root$ is extended to labeled {\btree s} by simply ignoring the extra labels. Also, $\lambda$ and $\oplus$ are extended to labeled {\btree s} in the obvious way: $$ \begin{tikzpicture}[ scale=0.45, baseline=(111.base) ] \style \path node [disc] (1) at ( 0, -1) {} node [disc] (11) at ( 0, -2) {} node [disc] (111) at (-.65, -3) {} node [disc] (112) at (0.65, -3) {}; \draw (1) node[right=2pt] {$(2,a)$} -- (11) node[right=2pt] {$(2,b)$} (11) -- (111) node[below=1pt] {$(1,c)$\;\;} (11) -- (112) node[below=1pt] {\;\;$(1,d)$}; \end{tikzpicture} = \lambda\Big((2,a),\!\! \begin{tikzpicture}[ scale=0.45, baseline=(111.base) ] \style \path node [disc] (11) at ( 0, -2) {} node [disc] (111) at (-0.65, -3) {} node [disc] (112) at ( 0.65, -3) {}; \draw (11) node[above=2pt] {$(2,b)$}-- (111) node[below=1pt] {$(1,c)$\;\;} (11) -- (112) node[below=1pt] {\;\;$(1,d)$}; \end{tikzpicture} \Big) = \lambda\Big((2,a),\! \begin{tikzpicture}[ scale=0.45, baseline=-2pt ] \style \node [disc] (r) at (0,1) {}; \node [disc] (1) at (0,0) {}; \draw (r) node[above=1pt] {\ensuremath{(1,b)}} -- (1) node[below=1pt] {\ensuremath{(1,c)}}; \end{tikzpicture} \!\!\oplus\!\! \begin{tikzpicture}[ scale=0.45, baseline=-2pt ] \style \node [disc] (r) at (0,1) {}; \node [disc] (1) at (0,0) {}; \draw (r) node[above=1pt] {\ensuremath{(1,b)}} -- (1) node[below=1pt] {\ensuremath{(1,d)}}; \end{tikzpicture} \Big) $$ Similarly, we extend $\gamma$ and $\obslash$: $$ \begin{tikzpicture}[ scale=0.45, baseline=(111.base) ] \style \path node [disc] (1) at ( 0, -1) {} node [disc] (11) at ( 0, -2) {} node [disc] (111) at (-.65, -3) {} node [disc] (112) at (0.65, -3) {}; \draw (1) node[right=2pt] {$(2,a)$} -- (11) node[right=2pt] {$(2,b)$} (11) -- (111) node[below=1pt] {$(1,c)$\;\;} (11) -- (112) node[below=1pt] {\;\;$(1,d)$}; \end{tikzpicture} = \gamma\Big((2,d), \begin{tikzpicture}[ scale=0.45, baseline=(111.base) ] \style \path node [disc] (1) at ( 0, -1) {} node [disc] (11) at ( 0, -2) {} node [disc] (111) at ( 0, -3) {}; \draw (1) node[right=2pt] {$(1,a)$} -- (11) node[right=2pt] {$(1,b)$} (11) -- (111) node[below=1pt] {$(1,c)$}; \end{tikzpicture}\, \Big) = \gamma\Big((2,d),\! \begin{tikzpicture}[ scale=0.45, baseline=-2pt ] \style \node [disc] (r) at (0,1) {}; \node [disc] (1) at (0,0) {}; \draw (r) node[above=1pt] {\ensuremath{(1,a)}} -- (1) node[below=1pt] {\ensuremath{(1,b)}}; \end{tikzpicture} \!\!\obslash\!\! \begin{tikzpicture}[ scale=0.45, baseline=-2pt ] \style \node [disc] (r) at (0,1) {}; \node [disc] (1) at (0,0) {}; \draw (r) node[above=1pt] {\ensuremath{(1,b)}} -- (1) node[below=1pt] {\ensuremath{(1,c)}}; \end{tikzpicture} \Big) $$ The involution $h$ is also easy to extend to \btree s: \begin{align*} h(1,x) &= (1,x); \\ h\lambda((i,x),t) &= \gamma((i,x),h(t));\\ h(u\oplus v) &= h(v)\obslash h(u). \end{align*} For instance, $$ \begin{tikzpicture}[ scale=0.45, baseline=(111.base) ] \style \node[font=\normalsize] at (-4,-2) {$t\;=$}; \path node [disc] (1) at ( 0, -1) {} node [disc] (11) at (-0.8, -2) {} node [disc] (12) at ( 0.8, -2) {} node [disc] (111) at (-0.8, -3) {} node [disc] (121) at ( 0.8, -3) {}; \draw (1) node[above=2pt] {$(2,a)$} -- (11) node[left =2pt] {$(1,b)$} (1) -- (12) node[right=2pt] {$(1,d)$} (11) -- (111) node[left =2pt] {$(1,c)$} (12) -- (121) node[right=2pt] {$(1,e)$}; \draw[->, semithick] (4,-2) -- (5,-2) node[font=\normalsize, midway, yshift=3mm] {$h$}; \path node [disc] (1) at ( 9.0, -1) {} node [disc] (11) at ( 8.2, -2) {} node [disc] (12) at ( 9.8, -2) {} node [disc] (121) at ( 9.0, -3) {} node [disc] (122) at (10.6, -3) {}; \draw (1) node[above=2pt] {$(2,e)$} -- (11) node[left =2pt] {$(1,d)$} (1) -- (12) node[right=2pt] {$(1,c)$} (12) -- (121) node[left =2pt] {$(1,b)$} (12) -- (122) node[right=2pt] {$(1,a)$}; \node[font=\normalsize] at (13.5,-2) {$=\;h(t)$}; \end{tikzpicture} $$ Let $\NODES(t)$ be the word obtained from preorder traversal of $t$. Also, denote by $w^r$ the reverse of the word $w$. For instance, with $t$ as above, we have $\NODES(t)=abcde$ and $\NODES(t)^r=edcba$. Let $\LEAVES t$ be the subword of $\NODES(t)$ whose letters are labels of leaves of $t$, and let $\INTERNAL t$ be the subword of $\NODES(t)^r$ whose letters are labels of internal nodes of $t$. Any labeled \mbox{\ensuremath{\beta(1,0)}-tree}\ can be written uniquely as a sum of indecomposable \btree s. If $t=\lambda((i_1,x_1),t_1)\oplus\dots\oplus\lambda((i_k,x_k),t_k) $ is so written, we let $\SUB t =(\,\NODES(t_1),\dots,\NODES(t_k)\,)$. Similarly, assuming that $t=\gamma((i_1,x_1),t_1)\obslash\dots\obslash\gamma((i_k,x_k),t_k)$ we let $\RSUB t = (\,\NODES(t_k)^r,\dots,\NODES(t_1)^r\,)$. The definition of the statistic $\ROOT t$ is a bit involved: $\ROOT t$ is a subword of $\LEAVES t$ of length $k=\root t$. More precisely, we build this word by starting at the root and greedily and recursively searching for $k$ leaves in its subtrees starting from the rightmost subtrees; also, we never search for more leaves in a subtree than the root label of that subtree. A more precise and formal definition can be found in the proof of Theorem~\ref{thm_h2}. Let $\RPATH t$ be the subword of $\NODES(v)^r$ whose letters are labels of the right path of $t$, excluding the leaf. With $t$ and $h(t)$ as in the above picture we have $$ \begin{array}{lclcl} \LEAVES t &=& \INTERNAL h(t) &=& ce \\ \INTERNAL t &=& \LEAVES h(t) &=& dba \\ \ROOT t &=& \RPATH h(t) &=& ce \\ \RPATH t &=& \ROOT h(t) &=& da \\ \SUB t &=& \RSUB h(t) &=& (bc,de) \\ \RSUB t &=& \SUB h(t) &=& (d,cba). \\ \end{array} $$ \begin{theorem}\label{thm_h2} On labeled {\btree s} with at least one edge, the involution $h$ sends the first tuple below to the second. $$ \begin{array}{lllllllll} ( &\LEAVES, &\INTERNAL, &\ROOT, &\RPATH, &\SUB, &\RSUB & ) \\ ( & \INTERNAL, &\LEAVES, &\RPATH, &\ROOT, &\RSUB, &\SUB & ) \end{array} $$ \end{theorem} \begin{proof} In terms of the recursive decomposition of labeled \btree s, we have $$ \begin{array}{lcl} \LEAVES\ (1,x) &=& x \\ \LEAVES \lambda((i,x),t) &=& \LEAVES t \\ \LEAVES (u\oplus v) &=& \LEAVES u \,\sqcup\, \LEAVES v \\[1.3ex] \INTERNAL\ (1,x) &=& x \\ \INTERNAL \gamma((i,x),t) &=& \INTERNAL t \\ \INTERNAL (u \obslash v) &=& \INTERNAL v \,\sqcup\, \INTERNAL u\\[1.3ex] \ROOT\ (1,x) &=& x \\ \ROOT \lambda((i,x),t) &=& \take_i(\ROOT t) \\ \ROOT (u\oplus v) &=& \ROOT u \,\sqcup\, \ROOT v \\[1.3ex] \RPATH\ (1,x) &=& x \\ \RPATH \gamma((i,x),t) &=& \take_i(\RPATH t) \\ \RPATH (u \obslash v) &=& \RPATH v \,\sqcup\, \RPATH u \\[1.3ex] \SUB\ (1,x) &=& \ensuremath{\epsilon} \\ \SUB \lambda((i,x),t) &=& (\NODES(t)) \\ \SUB (u \oplus v) &=& \SUB u \,\sqcup\, \SUB v \\[1.3ex] \RSUB\ (1,x) &=& \ensuremath{\epsilon} \\ \RSUB \gamma((i,x),t) &=& (\NODES(t)^r) \\ \RSUB (u \obslash v) &=& \RSUB v \,\sqcup\, \RSUB u \end{array} $$ where $u\/\sqcup\/v$ denotes the concatenation of $u$ and $v$, and $\take_i(a_1\dots a_n) = a_1\dots a_i$. Using induction and the definition of $h$ the result readily follows. \end{proof} \bibliographystyle{plain}
\section{Introduction} If two particles collide near the black hole horizon, the energy in their centre of mass (CM) frame can grow indefinitely. This interesting effect was discovered by Ba\~{n}ados, SIlk and West \cite{ban} and is called the BSW effect after the names of its authors and is under active study now. To achieve this effect, the collision by itself is insufficient. Additionally, one of particles has to have fine-tuned parameters, so that the some relationship between its energy and angular momentum or charge should hold. This somewhat restricts the realization of the BSW effect or requires special scenarios like collisions near circular orbits (that can occur in the accretion disc) \cite{circ}. The aim of the present work is point out that there exists one more class of objects - so-called quasblack holes (QBH) - for which the BSW effect can reveal itself. What is especially interesting is that for QBHs no fine-tuning is required at all, so the BSW effect is even more universal in this sense than one could think before. QBH is an object, roughly speaking, on the verge of collapsing and forming the event horizon but without actual forming the horizon. This includes a quite diverse set of physical objects: the self-gravitating monopole, so-called extremal dust and even a usual thin shell in the near-horizon limit (see \cite{qbh} and references therein for more strict definition and discussion of general properties). Especially interesting is the fact that QBHs realize the limiting transition from perfectly rotating fluid \cit {mein1} or a rapidly rotating disc \cite{disc} to the extremal black hole that can have important consequences in relativistic astrophysics. However, as the rotating (more realistic) case is more complicated, it looks reasonable to analyze the spherically symmetric space-time. It is expected to posses qualitatively similar features since rotation and charge have much in common in black hole physics.\ Meanwhile, it is much more easy for analysis. We do not touch upon here the issue of stability of QBHs which is a separate question. We only point out that at present, there are some preliminary results that lend support to the idea that due to the pressure, QBHs can be stable and correspond to frozen stars \cite{pres}. There is also additional motivation for studying the BSW effect in this context. There are works in which the analogue of the BSW effect is considered when the horizon is absent \cite{pj}, \cite{p}. The corresponding approach is essentially based on effects of self-gravitation of moving shells, so it looks more simple to reveal some features of the BSW effect without horizons in a more simple situation. In the present paper I restrict myself with the radial motion of charged particles in spherically symmetric metric that represents the counterpart of similar kind of the BSW effect for rotating black holes \cite{jl}.\ Meanwhile, the rotational analogue of QBHs also exists \cite{rot}, so the results can be extended to it in a straightforward manner. \section{BSW effect for black holes} Let us remind the main ingredients of the BSW effect for charged spherically symmetric black holes \cite{jl}. The metric can be written in the for \begin{equation} ds^{2}=-f(r)dt^{2}+\frac{dr^{2}}{A(r)}+r^{2}d\omega ^{2}\text{.} \label{m} \end{equation In the particular case of the Reissner-Nordstr\"{o}n (RN) metric $f=A=1 \frac{2M}{r}+\frac{Q^{2}}{r^{2}}$ where $M$ is the mass of the black hole, Q $ is its charge, the electric potential $\varphi =\frac{Q}{r}$. We use the systems of units in which the fundamental constants $G=c= h{\hskip-.2em}\llap{\protect\rule[1.1ex]{.325em}{.1ex}}{\hskip.2em =1. I consider, for definiteness, the extremal case, so for the RN black hole $M=Q=r_{+}$ where $r_{+}$ is the horizon radius, $f=\left( 1-\frac{r_{+ }{r}\right) ^{2}$. The equations of motion give us \begin{equation} mu^{0}=m\dot{t}=\frac{1}{f}(E-q\varphi )\text{,} \label{u0} \end{equation} \begin{equation} m^{2}\dot{r}^{2}=\frac{A}{f}(E-q\varphi )^{2}-m^{2}A\text{,} \label{ur} \end{equation where dot denotes derivative with respect to the proper time of a moving particle. Here, $m$ is the mass of a particle, $q$ is its electric charge, E $ is its conserved energy. Then, one can define the energy in the CM frame $E_{c.m.}$ according to E_{c.m.}^{2}=-P^{\mu }P_{\mu }$ where $P^{\mu }=p_{1}^{\mu }+p_{2}^{\mu }$ is the total momentum, $p^{\mu }$ are individual ones of particles, subscript $i$ enumerates particles' characteristics ($i=1,2$)$.$ Assuming for simplicity that $m_{1}=m_{2}=m$, one finds tha \begin{equation} \frac{E_{cm}^{2}}{2m^{2}}=1+\frac{X_{1}X_{2}-Z_{1}Z_{2}}{N^{2}m^{2}} \label{exz} \end{equation where \begin{equation} X_{i}=E_{i}-q_{i}\varphi \text{, }Z_{i}=\sqrt{X_{i}^{2}-m^{2}N^{2}} \label{xz} \end{equation and we introduce the lapse function, $f=N^{2}\,.$ Let particle 1 be critical and particle be usual. By definition, this means that $\left( X_{1}\right) _{+}=0$, hereafter subscript "+" denotes quantities calculated on the horizon. Then, \begin{equation} E_{1}=q_{1}\varphi (r_{+}) \label{cr} \end{equation and $\left( X_{2}\right) _{+}\neq 0$. If $\left( X_{1}\right) _{+}$ does not vanish exactly but is small in some sense (see below), we will call such a particle "near-critical". Near the horizon, we can use the Taylor expansion for the potential $\varphi \approx \varphi (r_{+})[1-a\frac{(r-r_{+})}{r_{+}}]+O((r-r_{+})^{2})$ and the lapse function $N\approx b\frac{(r-r_{+})}{r_{+}}$ where we took into account that for the extremal horizon, by definition, $f\sim (r-r_{+})^{2}$. In the RN case, the constants $a=b=1$. Then, if is easy to obtain from (\ref{xz}), (\ref{cr}) that for a critical particle \begin{equation} X_{1}\approx \frac{a}{b}E_{1}N\text{, } \end{equation near the horizon. Then, in the limit $f\rightarrow 0$ one can find tha \begin{equation} E_{cm}^{2}\approx \frac{2X_{2(H)}}{N}[\frac{a}{b}E_{1}-\sqrt{(\frac{a}{b )^{2}E_{1}^{2}-m^{2}}]\text{.} \end{equation} Thus the CM energy grows unbound that gives us the BSW effect. \section{Case of quasiblack holes} In recent years, a new physical object was singled out as a separate physical entity - so-called quasiblack holes. Roughly speaking, a quasiblack hole is a system on the threshold of formation of the event horizon when a system can approach this threshold as close as one likes but actually the horizon does not form. This includes a number of physical objects - self-gravitating Higgs magnetic monopoles \cite{lw1}, \cite{lw2}, extremal charged dust \cite{bon}, \cite{je} and even simple extremal shells \cite{qbh . More rigorous definition, brief review of existing realizations with extended list of references and discussion of general properties of quasblack holes can be found in \cite{qbh}, see also recent work \cite{rec}. To formulate the problem in mathematical terms, let us consider a compact body for which $N(r_{B})=\varepsilon $ is small on the boundary $r=r_{B}$. We assume that inside ($r<r_{B}$ \begin{equation} N=\varepsilon \tilde{N}(r) \end{equation where $\tilde{N}(r)\neq 0$, $\tilde{N}(r_{B})=1$ because of continuity of the metric on the boundary. Eq. (\ref{exz}) is still valid. Inside, one can introduce a new time coordinate according to $\tilde{t}=\varepsilon t$. Then, in the inner region the metric takes the form \begin{equation} ds^{2}=-\tilde{N}^{2}(r)d\tilde{t}^{2}+\frac{dr^{2}}{A(r)}+r^{2}d\omega ^{2 \text{.} \end{equation The small parameter $\varepsilon $ is eliminated, so $\tilde{t}$ is a natural measure of time in the inner region. Let particle 2 enter the inner region (say, from infinity) and collide there with particle 1. We assume that particle 2 was created \textit{inside} the boundary and has some finite energy $\tilde{E}$ measured with respect to the \textit{inner time }$\tilde{t}$. As the energy is a time component of the four-vector, $E=\tilde{E}\frac{\partial \tilde{t}}{\partial t}=\varepsilon \tilde{E}$. In a similar way, $\varphi =\varepsilon \tilde{\varphi}$. As a result, $X_{1}=\varepsilon \tilde{X}_{1}$ where $\tilde{X}_{1}$ does not contain a small parameter. It is clear from the formulas of the previous Section that it is smallness of $\left( X_{1}\right) _{+}$ that causes indefinite growth of $E_{c.m.}$. But this is just what happens now! Indeed, it follows from (\ref{exz}) tha \begin{equation} E_{cm}^{2}\approx \frac{2X_{2(H)}}{\varepsilon \tilde{N}^{2}}[\tilde{X}_{1} \sqrt{\tilde{X}_{1}^{2}-m^{2}\tilde{N}^{2}}]\text{.} \label{qe} \end{equation} Therefore, if $\varepsilon \ll 1,$ $E_{cm}^{2}$ can be made as large as one likes. \section{Discussion: comparison of two kinds of the BSW effect} The existence of the BSW effects for black holes is due to the horizon: the metric function $f=0$ at some point $r=r_{+}$. By contrast, in the QBH case, the metric function does not vanish at all. Instead, it remains small but nonzero in the entire region $r\leq r_{B}$. Correspondingly, the BSW effect can occur there not only near the would-be horizon (like near the horizon of black holes) but \textit{everywhere} inside. Actually, this is nontrivial consequence of mismatch between the time coordinates inside and outside that is important property of QBH \cite{qbh}. We saw in the preceding Section that $X_{1}\rightarrow 0$ for any typical particle created inside. In this sense, any such particle is near-critical. Meanwhile, particle 2 is usual, so the main condition of the BSW effect is satisfied. It is worth stressing that this condition is obtained without fine-tuning at all, so \textit{any} particle inside with a finite energy \tilde{E}$ is near-critical. As particle 2 is chosen to be usual, their collision leads to the BSW effect. It is also instructive to describe the situation in a kinematic language. For black holes, a critical particle approaches the horizon with some finite velocity $V_{1}<1$ measured by a static or stationary observer. Any usual particle has $V_{2}\rightarrow 1$ near the horizon. As a result, the relative velocity of two particles $w\rightarrow 1$, the corresponding Lorentz gamma factor indefinitely grows and $E_{c.m.}\rightarrow \infty $ (see \cite{k} for details). For radial motion in the metric (\ref{m}), $V=\frac{dl}{d\tau _{st}}$ where dl$ is the proper distance and $d\tau _{st}=\sqrt{f}dt$ is the proper time measured by a static observer in a given point. Using (\ref{m}), one finds \begin{equation} V=\frac{\left\vert \dot{r}\right\vert }{\sqrt{\dot{r}^{2}+A}}\text{,} \label{v} \end{equation In the inner region, $\tilde{N}$ $\ $and $\tilde{A}$ should be used here instead of $N$ and $A$. In the black hole case (say, for the RN metric), $A\sim f\rightarrow 0$ near the horizon. If particle is usual, $\dot{r}^{2}\neq 0$ due to the first term in the right hand side of (\ref{ur}), so $V\rightarrow 1$ for a usual particle. In a similar way, for the critical one, $\dot{r}^{2}\sim N^{2}\rightarrow 0$ and we have from (\ref{v}) that $V_{1}<1$. For QBHs the situation is somewhat different. The quantity $A$ is small near the boundary from the outside but is finite nonzero in the inner region (say, we can take Minkowski \ vacuum inside and the RN metric outside) \cit {qbh}. The quantity $N\sim \varepsilon $ is small inside. Let us consider the process from the viewpoint of an observer who resides inside the boundary $r_{B}$ and uses time $\tilde{t}$. Particle 1 created just inside with the finite energy $\tilde{E}$, has also finite $\dot{r}$, $\tilde{N}$, \tilde{A}$. As a result, $V_{1}<1$. From the other hand, for a generic particle 2 it follows from eq. (\ref{ur}) that $\left\vert \dot{r \right\vert \rightarrow \infty $ since $A$ is finite but $f\sim \varepsilon ^{2}$ is small in the inner region. Therefore, it follows from (\ref{v}) that $V_{2}\rightarrow 1$. Thus, the mechanisms that give rise to $V_{1}<1$, $V_{2}\rightarrow 1$ in both cases are different but the result is the same: the BSW effect due to the relative velocity approaching the speed of light. Thus, we obtained by different methods that particle 1 in this context is always near-critical, so due to the collision between a near-critical and usual particles the BSW effect occurs. The situation for which the BSW effect occurs can be summarized in the table. \begin{tabular}{|l|l|l|l|l|l|l|} \hline Particle & Origination & $E$ & $\tilde{E}$ & $V$ inside & $\left\vert \dot{r \right\vert $ inside & Type \\ \hline 1 & created inside & $\sim \varepsilon $ & finite & $<1$ & finite & near-critical \\ \hline 2 & coming from outside & finite & $\sim 1/\varepsilon $ & $\approx 1$ & \sim 1/\varepsilon $ & usual \\ \hline \end{tabular} It is worth mentioning that a quite different mechanism for the BSW effect for regular space-times without horizons was considered in \cite{pj}. It applies to pure radial motion of uncharged particles, required some special limitations on the velocities and angular momentums and leads to the dependence $E_{cm}^{2}\sim f^{-1}=N^{-2}$ whereas in our case E_{cm}^{2}\sim N^{-1}$ and the only physical requirement is the smallness of the parameter $\varepsilon $. Two reservations are in order. First, in the present paper, we neglected the size of particles and treated them as point-like objects. If one takes into account the size of falling objects, one should also take into account the tidal forces acting on them. For $\varepsilon \rightarrow 0$, the tidal forces for an observer falling inside the QBH grow unbound (see Sec. III 1 of \cite{qbh}). This prevents an object of a finite size to penetrate from the outside to inside. But for sufficiently small size and small but nonzero $\varepsilon $ one can neglect this effect. More precisely, one can obtain restriction similar to that obtained in \cite{cl} in another context. Indeed, it was shown in \cite{qbh} that typical value of the Riemann tensor in the free-falling frame has the order $R\sim \varepsilon ^{-2}M^{-2}$ where $M$ is a black hole mass. We require that, say, $R<\frac{1}{l_{0}^{2}} , where $l_{0}$ is a typical scale of an object which can withstand gravitational tidal forces. Then, taking for a typical black hole mass M\sim $ 10$^{5}cm$ in geometrical units and $l_{0}=10^{-8}$ cm is an atomic size one obtains that $\varepsilon >10^{-13}$ that is not restrictive at all. Second, if an object of a finite size moves inside the QBH with matter its trajectory is not geodesic. However, one can expect that this does not affect our results significantly even if corrections due to nonzero size are taken into account since (i) the energy of an incoming particle is much larger than that inside since it contains a small parameter $\varepsilon $ (see the Table), (ii) it was shown in \cite{gc} that the BSW effect retains its validity even for nongeodesic motion. It is worth also reminding that one of realization of a QBH is a simple vacuum shell in which case there is no matter inside at all. \section{Summary and conclusions} Thus we extended the class of objects which can be sources of the BSW effect.\ In addition to extremal \cite{ban} or nonextremal \cite{gp} black holes and naked singularities \cite{nk}, quasiblack holes are added and the mechanism of creating the BSW effect is analyzed. It turned out that the BSW effect may occur not only near the distinct surfaces but in the entire region. In so doing, there is no need to achieve fine-tuning by special selection of the particles for obtaining critical ones which are necessary for the BSW effect. Therefore, the results obtained look quite universal and can be useful in the analysis of physics of relativistic objects which are close to the stage of gravitational collapse. The possibility to accelerate particles is one more interesting property to QBHs in addition to those discussed in \cite{qbh}. In the present paper we restricted ourselves by radial motion in the background of spherically symmetric QBHs. Meanwhile, the rotating analogue of QBH appears naturally in astrophysics of rapidly rotating objects (see \cite{mein1} and references therein). Correspondingly, the BSW effect should occur also for rotating QBHs.
\section{Introduction} \label{intro} Hot, massive stars possess strong winds driven by line scattering of the star's intense ultra-violet (UV) radiation field. The first quantitative description of such line driving was given in the seminal paper by \citet[][CAK]{Castor75}, who assumed a smooth, steady-state outflow. But even though extensions of this theory \citep[e.g.,][]{Pauldrach86, Friend86} have had considerable success in explaining many global properties of OB-star winds, like the predicted mass-loss dependence on metallicity and the relation between the wind-momentum and the star's luminosity, it is nowadays clear that these winds are in fact both highly variable and structured on a broad range of temporal and spatial scales \citep[see][for comprehensive reviews]{Puls08, Sundqvist11b}. Linear stability analyses have shown \citep[][the last two hereafter ORI, ORII]{Macgregor79, Owocki84, Owocki85} that the line driving of these winds is subject to a very strong intrinsic instability, operating on small spatial scales. And subsequent numerical modeling of the non-linear evolution of this \textit{line-deshadowing instability} (LDI) has confirmed that the time-dependent wind indeed develops a highly inhomogeneous, `clumped', structure \citep{Owocki88, Feldmeier95, Dessart03}. Such structured LDI models provide a natural explanation for a number of observed phenomena in OB-stars, such as the soft X-ray emission and broad X-ray lines observed by orbiting telescopes like {\sc chandra} and {\sc xmm-newton} \citep{Feldmeier97, Berghofer97, Gudel09, Cohen10}, the extended regions of zero residual flux typically seen in saturated UV resonance lines \citep{Lucy83, Puls93, Sundqvist10}, and the migrating spectral sub-peaks superimposed on broad optical recombination lines \citep{Eversberg98, Dessart05b, Lepine08}. However, a multitude of independent observational studies also suggest the presence of clumps in inner wind regions close to the photosphere \citep{Eversberg98, Puls06, Sundqvist11, Najarro11, Cohen11, Bouret12}; this is \textit{not} reproduced by conservative, self-excited LDI models, which develop structure only away from the photospheric wind base \citep[at $r \ga 1.5 R_\star$,][]{Runacres02}. This result has led to a common perception that the LDI is not able to produce structure in inner wind regions, and so that some other process may be the main agent responsible for the overall features of wind clumping in OB-stars. But note that the strong damping of structure in the inner wind found in previous instability models was a direct consequence of a complete cancellation of the LDI by the counteracting line-drag effect \citep{Lucy84} at the stellar surface \citep[e.g.][hereafter OP96]{Owocki96}. But as already pointed out by ORII, if one accounts for limb-darkening of the stellar surface radiation, this cancellation should be incomplete and so lead to an unstable wind base. However, whereas the effect of such limb-darkening has been considered before in the context of steady-state, line-driven winds \citep{Cranmer95, Cure12}, it has never been explored in time-dependent simulations. This paper follows up on this early conjecture regarding the effect of limb-darkening on the LDI. In \S\ref{perturb} we perform an analytic linear stability analysis including simple Eddington limb-darkening of the photosphere. \S\ref{numerical} then describes our numerical, self-consistent, radiation-hydrodynamical modeling technique for studying the non-linear evolution of the competition between the LDI and the line-drag. \S\ref{results} examines to what extent the new models including limb-darkening, as well as a simple photospheric sound wave perturbation, can produce significant wind structure also in the inner wind. Finally, \S\ref{discussion} discusses these results, gives our conclusions, and outlines future work. \section{Analytic perturbation analysis including limb-darkening} \label{perturb} \begin{figure} \resizebox{\hsize}{!}{\includegraphics[]{fig1.pdf}} \caption{Relative damping effect of the line-drag on the ORI pure-absorption growth rate (eqn.~\ref{Eq:Om}) for uniformly bright (dashed lines) and limb-darkened (solid lines) stellar discs. Various approximations for the scattering source function as indicated in the figure, where `FI' denotes full angle integrations. Note how limb-darkened models are unstable also at the stellar surface ($\Omega_{\rm net}(R_\star) > 0$).} \label{Fig:net_growth} \end{figure} Let us first examine the effects of limb-darkening in the linear regime when perturbing the radiation force exerted by a single line. The instability arises from perturbations in the direct component of the line-force, which is proportional to the stellar core intensity $I_{\rm c}(\mu,r) = I_\star D(\mu,r)$. Here $I_\star$ sets the scale for the intensity and the flux-normalized disc function $D(\mu,r)$ accounts for the variation of intensity along local direction cosine $\mu$ at radius $r$. Previous analyses have ignored limb-darkening and simply assumed a uniformly bright stellar disc with $D=1$ for $\mu \ge \mu_\star \equiv \sqrt{1-(R_\star/r)^2}$ and $D=0$ for $\mu < \mu_\star$. For a line with frequency integrated mass absorption coefficient $\kappa$ and of `quality' $q \equiv v_{\rm th} \kappa/(\kappa_{\rm e} c)$ \citep{Gayley95}, we can write the perturbed direct component of the line acceleration in terms of an angle average of this core intensity (ORI, OP96) \begin{equation} \delta g_{\rm dir}(r) = \frac{4 \pi q \kappa_{\rm e}}{c} \langle \mu I_\star D(r,\mu) \, \delta b(r,\mu) \rangle, \label{Eq:dgdir} \end{equation} where $\delta b$ is the perturbed escape probability. For an optically thick line with Sobolev optical depth $\tau_\mu = \kappa \rho L_\mu \gg 1$, where $L_\mu = v_{\rm th}/(dv_{\rm n}/dn)$ is the Sobolev length in direction $\textbf{n}$, we find for perturbations on length scales below $L_\mu$ that $\delta b / \delta v \propto \mu/\tau_\mu$. Upon averaging, this leads to a strong instability with growth rate $\delta g_{\rm rad}/ \delta v \approx v/L_1$. Since this is a factor $v/v_{\rm th}$ larger than the wind expansion rate $dv/dr$, small initial velocity perturbations are strongly amplified within this linear theory, by $\approx v_\infty / v_{\rm th} \approx 100$ e-folds (ORI). This implies such small-scale perturbations will quickly reach non-linear amplitudes within pure-absorption, non-Sobolev wind models, as first demonstrated by \citet{Owocki88}. However, this strong de-shadowing instability can be counteracted by a `line-drag' effect \citep{Lucy84} associated with the force of the diffuse, scattered radiation field. Neglecting perturbations in the source function $S$, the perturbed diffuse force term is (ORII, OP96) \begin{equation} \delta g_{\rm diff}(r) = - \frac{4 \pi q \kappa_{\rm e}}{c} S(r) \langle \mu \, \delta b(r,\mu) \rangle, \label{Eq:dgdiff} \end{equation} where the minus sign signals the tendency of the line-drag to counteract the LDI. The similarity between eqs.~\ref{Eq:dgdir} and \ref{Eq:dgdiff} now allows us to write a very simple expression for the net relative reduction of the pure-absorption instability growth rate by the damping effect of this line-drag, \begin{equation} \Omega_{\rm net}(r) \equiv \frac{\delta g_{\rm dir} + \delta g_{\rm diff}}{\delta g_{\rm dir}} = 1 - \frac{ S(r) \langle \, \mu \, \delta b(r,\mu) \, \rangle} {\langle \, I_\star D(r,\mu) \, \mu \, \delta b(r,\mu) \, \rangle}. \label{Eq:Om} \end{equation} A key issue for evaluating eqn.~\ref{Eq:Om} lies in the computation of the assumed smooth source function $S$ for the unperturbed background flow. For a pure scattering line in a supersonic steady-state wind, this can be well approximated by the local Sobolev form \begin{equation} S(r) = \frac{ \langle \, I_\star D(r,\mu) b_{\rm Sob}(r,\mu) \, \rangle} {\langle \, b_{\rm Sob}(r,\mu) \, \rangle}, \label{Eq:sscat} \end{equation} where the Sobolev escape probability $b_{\rm Sob}(r,\mu) = (1-\rm e^{\it -\tau_\mu})/\tau_\mu$. For an optically thin line and a uniformly bright disc, the source function simply follows the dilution factor, $S(r)/I_\star = (1-\mu_\star)/2$. (Eqn.~\ref{Eq:S_ld} gives a corresponding expression including limb-darkening.) For a thick line, this diluted form has a further correction for the angle-averaged escape probability (see OP96). \subsection{Full angle integration} For an optically thick line, applying eqn.~\ref{Eq:sscat} in eqn.~\ref{Eq:Om} yields the asymptotic value $\Omega_{\rm net} \rightarrow 0.8$ as $r \rightarrow \infty$ (ORII). Such a modest 20\,\% reduction of the pure-absorption instability growth rate implies the wind still is extremely unstable in its outer parts. But closer to the star the damping is stronger; indeed, for a uniformly bright stellar disc with $S(R_\star) = I_\star/2$, the line-drag exactly \textit{cancels} the de-shadowing instability at the stellar surface, i.e. $\Omega_{\rm net} \rightarrow 0$ when $r \rightarrow R_\star$. This holds for both optically thick and thin scattering source functions, and leads to marginal stabilization of the wind base and to a later onset of wind structure than typically indicated by observations. However, noting that this exact cancellation occurs only because $S(R_\star) = I_\star/2$, let us now examine the effect of photospheric limb-darkening on this net growth rate. For a simple Eddington (linear) limb-darkening law (eqn.~\ref{Eq:ld}), analytic evaluation of eqn.~\ref{Eq:Om} gives $\Omega_{\rm net}(R_\star) = 0.06$ and $\Omega_{\rm net}(R_\star) = 0.22$ for, respectively, an optically thick and thin source function. This means that in such limb-darkened models the line-drag no longer exactly cancels the LDI at the stellar surface and thus that also the wind base is unstable. \subsection{Two-stream approximation} In a numerical wind model aiming to simulate the non-linear evolution of the LDI, it is very computationally expensive to carry out elaborate angle integrations at each radial grid-pint at each time-step. But testing has shown that a simple one-ray quadrature using a ray parameter $y \equiv (p/R_\star)^2 = 0.5$, with impact parameter $p \equiv r \sqrt{1-\mu^2}$, actually approximates rather well the full-angle integrated line force (see Appendix A), as long as separate accounts are taken for inward and outward rays in the diffuse force computations and a sphericity correction factor is applied (OP96). For this two-stream approximation, we find \begin{equation} \Omega_{\rm net}(r) = 1 - \frac{S(r)}{I_\star} \frac{2 (r/R_\star)^2}{D(\mu_y)}, \label{Eq:Om_ts} \end{equation} where $\mu_{\rm y} \equiv \sqrt{1- y (R_\star/r)^2}$ is the local direction cosine of the ray at radius $r$. Assuming an optically thin source function, this again results in a zero growth rate at the stellar surface for a uniform disc, for which eqn.~\ref{Eq:Om_ts} actually simplifies to $\Omega_{\rm net}(r) = \mu_\star/(1+\mu_\star)$ \citep[see also][]{Owocki99}. By contrast, for a limb-darkened disc (again with an optically thin source function), we find $\Omega(R_\star) = 0.15$, which is intermediate between the thin and thick results for full angle integration. Fig.~\ref{Fig:net_growth} plots $\Omega_{\rm net}(r)$ over $r/R_\star = 1 - 1.5$ for the cases discussed above, assuming a smooth velocity law $v = (1-R_\star/r)^\beta$ with $\beta =0.8$ \citep{Pauldrach86}. Note that over this inner wind range, the two-stream approximation can again be viewed as an intermediate case between the fully angle-integrated cases with optically thick and thin source functions. Most significantly, since the pure-absorption growth rate $\delta g_{\rm dir}$ is of order 100 times the wind expansion rate, even the base-value of $\sim$\,20\,\% of this rate found in this paper for limb-darkened models implies an absolute growth that is still substantially faster, by a factor $\sim$\,20, than the wind expansion. This suggests that non-linear structure can now develop close to the wind base, as we next demonstrate. \section{Numerical simulations of the time-dependent wind} \label{numerical} To follow the non-linear evolution of such instability-generated wind structure, we must numerically integrate the radiation-hydrodynamical conservation equations of mass, momentum, and energy. Table~\ref{Tab:params} summarizes the stellar and wind parameters adopted in the models, which correspond to a typical early O-supergiant in the Galaxy. The basic simulation method used here has been extensively described in \citet{Owocki88}, OP96, \citet{Runacres02}; this section recapitulates key assumptions and describes new features. \begin{table} \centering \caption{Summary of stellar and wind parameters} \begin{tabular}{p{2.8cm}ll} \hline \hline Name & Parameter & Value \\ \hline Stellar luminosity & $L_\star$ & $8\,\times\,10^{5}\,\rm L_\odot$ \\ Stellar mass & $M_\star$ & 50 \,$\rm M_\odot$ \\ Stellar radius & $R_\star$ & 20 \,$\rm R_\odot$ \\ Wind floor & & \\ temperature & $T_{\rm w}$ & 40\,000\,K \\ Initial steady-state & & \\ \ - terminal speed & $v_\infty$ & 2000\,km/s \\ \ - mass-loss rate & $\dot{M}$ & $2.1\,\times\,10^{-6}\,\rm M_\odot/yr$ \\ CAK exponent & $\alpha$ & 0.65 \\ Line-strength & & \\ \ - normalization & $\bar{Q}$ & 2000 \\ \ - cut-off & $Q_{\rm max}$ & 0.01$\bar{Q}$ \\ Ratio of ion thermal & & \\ speed to sound speed & $v_{\rm th}/a$ & 0.28 \\ Electron scattering & & \\ opacity & $\kappa_{\rm e}$ & 0.34\,$\rm cm^2/g$ \\ \\ \hline \end{tabular} \label{Tab:params} \end{table} \subsection{Radiative cooling and radiative driving} We solve the spherically symmetric conservation equations using the numerical hydrodynamics code VH-1 (developed by J.~Blondin et al.). Radiative cooling of shock-heated gas is accounted for in the energy equation following \citet{Runacres02}. This method mimics the effect of photoionization heating from the star's intense UV radiation field by simply demanding that the gas never cools below a certain floor temperature, set here to 40\,000\,K (roughly the star's effective temperature). A central challenge in these simulations is to compute the line-driving in a highly structured, time-dependent wind with a non-monotonic velocity. This requires a non-local integration of the line-transport within each time-step of the simulation, in order to capture the instability near and below the Sobolev length. To allow for this objective, we adopt the smooth source function \citep[SSF,][]{Owocki91b} method described extensively in OP96. As in CAK, we assume a line-number distribution that is a power-law in line-strength $q$, but now (as in OP96) exponentially truncated at a maximum strength $Q_{\rm max}$, \begin{equation} \frac{dN}{dq} = \frac{1}{\Gamma(\alpha) \bar{Q}} \left( \frac{q}{\bar{Q}} \right)^{\alpha-2} e^{-q/Q_{\rm max}}, \label{Eq:powerlaw} \end{equation} with $\Gamma(\alpha)$ the complete gamma function. Here $\alpha$ is the CAK power-law index, which can be physically interpreted as the ratio of the line force due to optically thick lines to the total line force, and $\bar{Q}$ is a line-strength normalization constant, which can be interpreted as the ratio of the total line force to the electron scattering force in the case that all lines were optically thin\footnote{Note that we have recast the line force using the $\bar{Q}$ notation of \citet{Gayley95} rather than the $\kappa_0$ notation of OP96. $\bar{Q}$ has the advantage of being a dimensionless measure of line-strength that is independent of the thermal speed. The relation between the two parameter formulations is $\bar{Q} = \kappa_0 v_{\rm th} \Gamma(\alpha)^{1/(1-\alpha)}/(c \kappa_{\rm e})$.}. For typical O-supergiant conditions at solar metallicity, $Q_{\rm max} \approx \bar{Q} \approx 2000$ \citep{Gayley95, Puls00}. In practice, keeping the nonlinear amplitude of the instability from exceeding the limitations of the numerical scheme requires a significantly smaller cut-off \citep{Owocki88}. First test-simulations \citep[see also][]{Runacres02} that increase $Q_{\rm max}$ to more physically realistic values indeed display stronger structure in wind regions where the instability is fully grown. But this higher $Q_{\rm max}$ also leads to spurious spikes in the velocity field, likely numerical artefacts caused by the limiting grid resolution. The full implications of the line-strength cut-off for wind structure is yet to be determined, and should be carefully examined in a detailed parameter study. However, we defer that to future work, since the test-simulations here also show that increasing the cut-off does not significantly affect the main focus of this paper, namely the \textit{onset} of non-linear structure. All previous time-dependent wind simulations have assumed the stellar continuum radiation can be described by a uniformly bright stellar disc. But as discussed in \S\ref{perturb}, introducing photospheric limb-darkening breaks the marginally stable properties of the wind base and so may lead to structure formation also in the innermost wind. As also noted in \S\ref{perturb}, for computational reasons we use a simple one-ray angle quadrature for both the direct and diffuse force components in all numerical models (see Appendix A). Incorporating Eddington limb-darkening then typically increases the direct line force by a modest amount, $\sim 3\,\%$, whereas the scattering source function (and thus the diffuse line force) at the stellar surface \textit{decreases} by $\sim 13\,\%$. This is the essential reason there is now a net instability at the wind base. \subsection{Boundary conditions} Each instability simulation evolves a smooth and relaxed Sobolev wind model. The lower boundary is set as in \citet{Runacres02}, by fixing the density and temperature. Previous simulations have assumed the lowermost grid-point to be located at $r = R_\star$, the photospheric radius where the continuum optical depth is unity, with a density $\rho_0$ tuned to be approximately 5-10 times higher than at the sonic point\footnote{Using a higher value leads to long-lived oscillations at the Lamb frequency; using lower values risk choking the self-regulated mass loss of the line-driven wind \citep{Owocki88, Owocki99}.}. However, a simple ratio estimate of the steady-state sonic point density $\rho_{\rm a} = \dot{M}/(4 \pi a r^2)$ to the typical photospheric density $\rho_\star \approx 1/(H \kappa_{\rm e})$, with scale height $H = a^2/g$, yields for the stellar and wind parameters in table~\ref{Tab:params} a much lower value, $\rho_{\rm a}/\rho_\star < 0.01$. This order-of-magnitude estimate is confirmed by more detailed, line-blanketed NLTE calculations of steady-state, unified (photosphere+wind) model atmospheres of typical O supergiants like $\zeta$ Pup \citep[using the {\sc fastwind} code,][]{Puls05}. In such models, we find that the sonic point is located $\sim$\,1-2\,\% above the stellar surface, defined by $R_\star= r(\tau_{\rm Ross}=2/3)$, and that the corresponding density contrast is $\rho_{\rm a}/\rho_\star \approx 0.01$. The time-dependent simulations reported here thus assume $r \approx 1.01 R_\star$ at the sonic point, which tends to shift inward the stellar surface somewhat as compared to earlier models. And since the stability properties of the lower wind are so sensitive to the specific value of the source function, which close to the wind base decreases very rapidly with radii, accounting for this shift can actually be quite significant (particularly for lower-density winds). This adds to the effect of limb-darkening in making the region near the sonic point have a net instability. Moreover, if one accounts also for the likelihood that the stellar photosphere is not perfectly steady, but rather itself has a level of variability, this can further seed the growth of unstable structure in the overlying wind. Some previous simulations have explicitly perturbed the lower boundary by either sound waves or turbulence \citep[e.g.,][]{Feldmeier97}, and indeed such models tend to show structure somewhat closer to the wind base as compared to models with self-excited structure (compare \citealt{Runacres02} with \citealt{Puls93} and \citealt{Sundqvist11}). To study the influence of such perturbations in the presence of limb-darkening, the simulations here introduce a lower boundary sound wave of amplitude $\delta \rho/\rho_0 = 0.1$ and period 4\,000\,s. \section{Numerical simulation results} \label{results} \subsection{Basic structure properties} \begin{figure*} \begin{minipage}{5.8cm} \resizebox{\hsize}{!}{\includegraphics[]{fig2a.jpg}} \end{minipage} \begin{minipage}{5.8cm} \resizebox{\hsize}{!}{\includegraphics[]{fig2b.pdf}} \end{minipage} \begin{minipage}{5.8cm} \resizebox{\hsize}{!}{\includegraphics[]{fig2c.pdf}} \end{minipage} \caption{The left panel shows density and velocity contour plots of the time-evolution of a wind model with self-excited structure assuming a uniformly bright stellar surface. The middle and right panels show density and velocity of a single snapshot from the same model, taken $\sim 150 \, \rm ksec$ after initiation, as function of radius (middle) and a Lagrangian mass coordinate (right). The dashed red line is the smooth start model.} \label{Fig:basic} \end{figure*} Let us first review the overall properties of numerical simulations that follow the non-linear evolution of the competition between the LDI and the line-drag \citep{Owocki99, Runacres02, Dessart03}. The simulation displayed in Fig.~\ref{Fig:basic} illustrates the formation of high-speed rarefactions that steepen into strong reverse shocks, which compress most of the wind material into dense and spatially narrow `clumps' (or really shells in these 1-D simulations). This characteristic structure can be alternatively illustrated by plotting density and velocity against a Lagrangian mass coordinate tracking individual fluid elements \citep[as defined by][their eqn. 14]{Owocki99}; the rightmost panel of Fig.~\ref{Fig:basic} clearly shows how the high-speed flow consists of very rarefied gas and how most of the wind mass is concentrated into dense clumps. This model is computed assuming a uniformly bright stellar disc and without explicitly perturbing the lower boundary. The structure that evolves from the smooth wind at $t = 0$ in the left panel of Fig.~\ref{Fig:basic} is ``self-excited'', and arises from back-scattering of radiation from outer wind structure seeding small variations closer to the wind base, which are then subsequently amplified by the instability \citep{Owocki99}. Note how the line-drag effect discussed in \S\ref{perturb} greatly suppresses instability growth close to the wind base; indeed, structure starts to develop only at $r \approx 1.5 R_\star$ in this simulation, at odds with observations which typically indicate the presence of dense clumps much closer to the stellar surface (see \S\ref{intro}). The next section examines to what extent including limb-darkening and photospheric perturbations into the simulations can induce wind structure also at such low radii. \subsection{Clumping in the inner wind} \begin{figure*} \begin{minipage}{8.5cm} \resizebox{\hsize}{!}{\includegraphics[]{fig3a.pdf}} \end{minipage} \begin{minipage}{8.5cm} \resizebox{\hsize}{!}{\includegraphics[]{fig3b.pdf}} \end{minipage} \caption{Simulated clumping factors $f_{\rm cl}(r)$. The left panel focuses on the inner wind regions $r/R_\star = 1 - 2$, whereas the right panel shows the wind all the way out to $r/R_\star = 10$ on a logarithmic abscissa. Blue dashed-dotted lines compare the simulations to the clumping law inferred for $\zeta$ Pup by \citet{Najarro11}, see text. `LD' indicates whether limb-darkening is accounted for in the models.} \label{Fig:fcl} \end{figure*} \begin{figure*} \begin{minipage}{8.5cm} \resizebox{\hsize}{!}{\includegraphics[]{fig4a.jpg}} \end{minipage} \begin{minipage}{8.5cm} \resizebox{\hsize}{!}{\includegraphics[]{fig4b.jpg}} \end{minipage} \caption{Inner wind time evolutions of a simulation without limb-darkening and photospheric perturbations (left) and one including both effects (right).} \label{Fig:time_ev} \end{figure*} Let us characterize the wind density structure in terms of a clumping factor \begin{equation} f_{\rm cl} = \langle \rho^2 \rangle / \langle \rho \rangle^2, \end{equation} where the angle brackets here denote time averaging. Fig.~\ref{Fig:fcl} plots the radial variation of such clumping factors for the four simulation cases considered, i.e. for uniform-disc and limb-darkened models with and without explicit base perturbations, computed by averaging over each simulation time-step between $t=100-300 \, \rm ksec$, where the simulations have become insensitive to the initial conditions. The right panel shows $f_{\rm cl}$ over the full simulation range $r/R_\star = 1-10$, whereas the left panel focuses on the crucial inner wind regions where limb-darkening and base perturbations alter the onset of wind structure. The blue dashed curve in both panels show, for comparison, the empirically inferred clumping factor\footnote{In such empirical work, $f_{\rm cl}$ is typically defined using volume-averaging rather than time-averaging, but in the stochastic medium here the two should be effectively interchangeable. If one neglects the small inter-clump density \citep[but see][]{Zsargo08, Sundqvist10, Surlan12}, this clumping factor is equal to the inverse of the clump volume filling factor, $f_{\rm cl} = 1/f_{\rm vol}$.} for $\zeta$ Pup based on the comprehensive multi-diagnostic study by \citet{Najarro11}. The left panel illustrates that both perturbations and limb-darkening shift inward the onset of clumping as compared to the standard model. Much as anticipated in the linear stability analysis, limb-darkening increases the net growth rate, while the perturbation provides a seed which the instability can amplify to form the clumped structure. The right panel shows that this stronger structure persists to several stellar radii from the wind base. But note that our simulation volume here only extends to $ \approx 10 R_\star$; the analysis of \citet{Runacres02} suggests that such strong structure will eventually dissipate and tend to settle at $f_{\rm cl} \approx 4$ in the outermost, radio emitting wind. We emphasize here that this initial exploration study only attempts to outline the general importance of limb-darkening and perturbations for the onset of LDI generated wind structure. Thus we do not aim for a perfect match to the empirical $\zeta$ Pup clumping law throughout the wind; quantitative details in the predicted clumping factors will depend on the exact treatment of limb-darkening and the source function, the nature and strength of the applied base perturbations, and the level of lateral fragmentation of clumps in multi-dimensional LDI simulations \citep{Dessart03, Dessart05}, as further discussed in \S\ref{discussion}. Nonetheless, the new overall agreement between simulations and observations in the inner wind is encouraging, and suggests that the standard line-deshadowing instability (possibly seeded by some modest photospheric perturbations) is indeed fully capable of generating significant structure also near the wind base. \subsection{Time-evolution with and without limb-darkening and base perturbations} Fig.~\ref{Fig:time_ev} compares the time evolution of velocity and density in the inner wind for the model including limb-darkening and base perturbations (right) against the model ignoring both effects (left). Following the brief adjustment to initial conditions, the unperturbed model settles to a state with onset of intrinsic structure and variability at $r \approx 1.5 R_\star$. In contrast, the perturbed, limb-darkened model shows a much earlier onset that varies from a maximum $r \approx 1.15 R_\star$ all the way down to the photospheric wind base. The characteristic time-scale of this variation, $\sim$\,50\,ksec, is much longer than the perturbation period of $\sim$\,4\,ksec. Our preliminary analysis suggests that this long-term variation is likely related to the intrinsic variations found in the ``pure-absorption'' models of \citet{Poe90}. As discussed there, these variations stem from a degeneracy of the CAK-like steady-state solutions in the regions near the wind sonic point. As shown in Fig.~7b of \citet{Poe90}, this leads to a wind-speed oscillation with period $\sim$\,50\,ksec. For slightly different conditions, the oscillating flow can actually even stagnate and re-accrete onto the star \citep[see Fig.~1 in][]{Owocki91}. Remarkably, in the standard SSF model the introduction of the diffuse force component eliminates this solution degeneracy, and so suppresses the oscillatory behaviour. But the additional introduction here of limb-darkening now enhances the direct pure-absorption component of the line-force, and reduces the stabilizing diffuse component. This evidently allows the degeneracy to reappear, and so again leads to slow variations in the wind solution. The full consequences of this effect for wind structure and variability (for example, as a possible origin of discrete absorption components, \citealt{Howarth89}) must await multi-dimensional LDI wind models \citep{Dessart03, Dessart05}. Similarly, implications for wind initiation must await a future, more complete analysis of the dynamical role of diffuse radiation near the sub-sonic wind base \citep[see, e.g.,][]{Owocki99}. Note though, that in the current models the mass flux initiated at the wind base is still quite close to the initial steady state derived from standard CAK Sobolev theory. \section{discussion, conclusions, and future work} \label{discussion} The central result of this paper is that photospheric limb-darkening in instability wind models leads to structure growth closer to the wind base than in previous models assuming a uniformly bright stellar disc. We demonstrate this both analytically in a linear perturbation analysis, and numerically using self-consistent, time-dependent radiation-hydrodynamical wind simulations. Particularly when combined with perturbations of the lower boundary, such limb-darkened models reproduce well the early onset of clumping typically inferred from observations of O-star winds. The analysis here uses a simple sound wave to examine the general effects of lower boundary perturbations for the onset of LDI-generated wind structure. Physically, photospheric perturbations could, for example, originate from non-radial pulsations, or even from the thin sub-surface convective layer associated with the iron opacity peak \citep{Cantiello09}. Moreover, as in previous models, the mass flux at the wind base in these time-dependent simulations follows quite closely that of the initial steady state derived from standard CAK theory, This suggests the early onset of wind structure found here does not directly alter the average mass-loss rate of the star. But note in this respect that the resulting clumped structure will modify e.g. the wind ionization balance, and so \textit{indirectly} affect the line-force. Such feedback effects are not included in the simulations presented here, but may impact theoretical mass-loss rates derived from steady-state models \citep{Muijres11}. For simplicity the simulations here assume spherical symmetry; first 2-D models by \citet{Dessart03, Dessart05} suggest somewhat lower clumping factors than in comparable 1-D models, by a factor of approximately two. Future work should carry out such multi-D LDI simulations including also the effects of limb-darkening and base perturbations. Finally, an unanticipated result of this study is the re-appearance of the solution degeneracy and associated slow global wind variations found in the pure-absorption models of \citet{Poe90}. This arises from the change in the relative strength of the diffuse vs. direct force components in the transonic wind region of limb-darkened simulations. Determining the broad implications of these effects for wind structure and variability will require a careful analysis of the diffuse radiation in this wind initiation region around the sonic point. Thus, although standard CAK wind models focus solely on the direct radiation from the star, this again \citep[see also, e.g.,][]{Owocki99} emphasizes the subtle role of diffuse scattered radiation in the dynamics of line-driven winds. \section*{Acknowledgments} This work was supported in part by NASA ATP grant NNX11AC40G. We would also like to thank the referee for useful comments on the manuscript. \newpage
\section{\protect\bigskip Introduction} Harmonic maps have been studied first by J. Eells and J.H.Sampson in the sixties and since then many articles have appeared ( see \cite{6}, \cite{12 , \cite{16}, \cite{19}, \cite{20}, \cite{24}) to cite a few of them. Extensions to the notions of $p$-harmonic, biharmonic, $F$-harmonic and $f -harmonic maps were introduced and similar research has been carried out (see \cite{1}, \cite{2}, \cite{3}, \cite{7}, \cite{15}, \cite{18}, \cite{21 , \cite{23}). Harmonic maps were applied to broad areas in sciences and engineering including the robot mechanics ( see \cite{5}, \cite{8}, \cite{9} )$.$The concept of $F$- harmonic maps unifies the notions of harmonic maps, p$-harmonic maps, minimal hypersurfaces. An important tool for studying stability of stability of $F$ harmonic maps is the stress-energy tensor. In this paper for a $C^{2}$-function $F:\left[ 0,+\infty \right[ \rightarrow \left[ 0,+\infty \right[ $ such that $F^{\prime }(t)>0$ on $t\in \left] 0,+\infty \right[ $, we look for sufficient conditions which present $F -harmonic maps into spheres as global maxima of the energy functional. Our result extends similar results obtained in \cite{17} and \cite{18} for harmonic and $p$-harmonic maps. Let $(M,g)$ and $S^{n\text{ }}$be, respectively, a compact Riemannian manifold of dimension $m\geq 2$ and $\ $the unit $n$-dimensional Euclidean sphere with $n\geq 2$ endowed with the canonical metric $can$ induced by the inner product of $R^{n+1}$. For a $C^{1}$- application $\phi :(M,g)\longrightarrow (S^{n},can)$, we define the $F$-energy functional by \begin{equation*} E_{F}(\phi )=\int_{M}F\left( \frac{\left\vert d\phi \right\vert ^{2}}{2 \right) dv_{g}\text{,} \end{equation* where $\frac{\left\vert d\phi \right\vert ^{2}}{2}$ denotes the energy density given by \begin{equation*} \frac{\left\vert d\phi \right\vert ^{2}}{2}=\frac{1}{2}\sum_{i=1}^{m}\lef \vert d\phi (e_{i})\right\vert ^{2} \end{equation* and where $\left\{ e_{i}\right\} $ is an orthonormal basis on the tangent space $T_{x}M$ and $dv_{g}$ is the Riemannian measure associated to $g$ on M $. Let $\phi ^{-1}TS^{n}$and $\Gamma \left( \phi ^{-1}TS^{n}\right) $ be, respectively, the pullback vector fiber bundle of $TS^{n}$ and the space of sections on $\phi ^{-1}TS^{n}$. Denote by $\nabla ^{M}$, $\nabla ^{S^{n}} and $\nabla $, respectively, the Levi-Civita connections on: $TM$, $T$ S^{n} $ and $\phi ^{-1}TS^{n}$. Recall that $\nabla $ is defined by \ \begin{equation*} \nabla _{X}Y=\nabla _{\phi _{\ast }X}^{S^{n}}Y \end{equation* where $X\in TM$ and $Y\in \Gamma \left( \phi ^{-1}TS^{n}\right) $. Let $v$ be a vector field on $S^{n}$ and denote by $\left( \gamma _{t}^{v}\right) _{t}$ the flow of diffeomorphisms induced by $v$ on $S^{n}$ i.e. \begin{equation*} \gamma _{0}^{v}=id\ \text{, \ \ }\frac{d}{dt}\gamma _{t}^{v}{}_{t=0}=v\left( \gamma _{t}^{v}\right) \text{.} \end{equation* Denote by $\phi _{t}=\gamma _{t}^{v}o\phi $ the flow generated by $v$ along the map $\phi $. The first variation formula of $E_{F}(\phi )$ is given b \begin{equation*} \frac{d}{dt}E_{F}(\phi _{t})\mid _{t=0}=\int_{M}F^{\prime }\left( \frac \left\vert d\phi _{t}\right\vert ^{2}}{2}\right) \left\langle \nabla _{\partial t}d\phi _{t},d\phi _{t}\right\rangle \left\vert _{t=0}\right. dv_{g} \end{equation* \begin{equation*} =-\int_{M}\left\langle v,\tau _{F}(\phi )\right\rangle dv_{g} \end{equation* where $\tau _{F}(\phi )=trace_{g}\nabla \left( F^{\prime }\left( \frac \left\vert d\phi \right\vert ^{2}}{2}\right) d\phi \right) $ is the $F -tension. \begin{definition} $\phi $ is said $F$-harmonic if and only if $\tau _{F}(\phi )=0$ i.e. $\phi $ is a critical point of $\ $the $F$-energy functional $E_{F}$. \end{definition} Let $v\in \mathbb{R} ^{n+1}$ and set $\bar{v}(y)=v-\left\langle v,y\right\rangle y$ \ for any y\in S^{n}$. It is known that $\bar{v}$ is a conformal vector field on S^{n} $ i.e. $\left( \gamma _{t}^{v}\right) ^{\ast }can=\alpha _{t}^{2}can$ where $\left( \gamma _{t}^{v}\right) _{t}$ denotes the flow induced by the vector field $\bar{v}$. The expression of $\alpha _{t}$ is given in \cite{17} by \begin{equation} \alpha _{t}=\frac{\left\vert v\right\vert }{\left\vert v\right\vert cht+\phi _{v}sht}\text{.} \label{1} \end{equation where $\phi _{v}\left( x\right) =\left\langle v,\phi \left( x\right) \right\rangle $ and $\left\langle .,.\right\rangle $ the inner product on the Euclidean space \mathbb{R} ^{n+1}$. Denote by $\pounds (\phi )$ the subspace of $\Gamma (\phi ^{-1}TS^{n})$ given by \begin{equation*} \pounds (\phi )=\left\{ \bar{v}\circ \phi ,v\in \mathbb{R} ^{n+1}\right\} \text{.} \end{equation* Obviously, if $\phi $ is not constant, $\pounds (\phi )$ is of dimension n+1 $. \section{ $F$-harmonic maps as global maxima} For any $\overline{v}\in \pounds (\phi )$, we denote by $\left( \gamma _{t}^{v}\right) _{t\in R}$ the one parameter group of conformal diffeomorphisms on $S^{n}$ induced by the vector $\bar{v}$. For\ a $C^{2} -function $F:\left[ 0,+\infty \right[ \rightarrow \left[ 0,+\infty \right[ $ such that $F^{\prime }(t)>0$ in $\left] 0,+\infty \right[ $. Now we introduce the following tensor field \begin{equation*} S_{g}^{F}\left( \phi \right) =F^{\prime }\left( \frac{\left\vert d\phi \right\vert ^{2}}{2}\right) \frac{\left\vert d\phi \right\vert ^{2}}{2}g \end{equation* \begin{equation*} -\left[ F^{\prime }\left( \frac{\left\vert d\phi \right\vert ^{2}}{2}\right) +\frac{\left\vert d\phi \right\vert ^{2}}{2}F^{\prime \prime }\left( \frac \left\vert d\phi \right\vert ^{2}}{2}\right) \right] \phi ^{\ast }can. \end{equation* For $x\in M$, we set \begin{equation*} S_{g}^{o,F}(\phi )\left( x\right) =\inf \left\{ S_{g,v}^{F}(\phi )(X,X)\text , }X\in T_{x}M\text{ such that }g(X,X)=1\right\} \text{.} \end{equation* The tensor field $S_{g}^{F}(\phi )\left( x\right) $ will be said positive ( resp. positive defined) at $x$ if $\ S_{g}^{o,F}(\phi )\left( x\right) \geq 0 $ (resp.$\ \ S_{g}^{o,F}(\phi )\left( x\right) >0$). The tensor field S_{g}^{F}(\phi )$ will be called the $F$ stress-energy tensor of $\phi $. The tensor field $S_{g}^{F}(\phi )$ is different from the one defined by Ara given by $S_{F}\left( \phi \right) =F^{{}}\left( \frac{\left\vert d\phi \right\vert ^{2}}{2}\right) g-F^{^{\prime }}\left( \frac{\left\vert d\phi \right\vert ^{2}}{2}\right) \phi ^{\ast }can$, but $S_{g}^{F}(\phi )$ is more suitable for our case. \begin{example} For $F(t)=\frac{1}{p}\left( 2t\right) ^{\frac{p}{2}}$, with $p=2$ or $p\geq 4 $, $S_{g}^{p}\left( \phi \right) $ is the stress-energy tensor introduced, respectively, by Eells and Lemaire for $p=2$ \cite{12} and modulo a multiplied positive constant by El Soufi for $p\geq 4$ \cite{16}, so we may call $S_{g}^{F}(\phi )$ the stress-energy tensor of $\phi $. \end{example} Indeed if $F(t)=t$ then $F\prime (t)=1$, $F^{\prime \prime }(t)=0$ an \begin{equation*} S_{g}^{F}\left( \phi \right) =\frac{\left\vert d\phi \right\vert ^{2}}{2 g-\phi ^{\ast }can. \end{equation* In the case $F(t)=\frac{1}{p}\left( 2t\right) ^{\frac{p}{2}}$, with $p\geq 4 , $F\prime (t)=(2t)^{\frac{p}{2}-1}$, $\ F^{\prime \prime }(t)=\left( p-2\right) \left( 2t\right) ^{\frac{p}{2}-2}$ an \begin{equation*} S_{g}^{F}\left( \phi \right) =\frac{1}{2}\left\vert d\phi \right\vert ^{p}g \frac{p}{2}\left\vert d\phi \right\vert ^{p-2}\phi ^{\ast }can=\frac{p}{2 \left( \frac{1}{p}\left\vert d\phi \right\vert ^{p}g-\left\vert d\phi \right\vert ^{p-2}\phi ^{\ast }can\right) \end{equation* The function $F$ is called \textit{admissible} if it satisfies \begin{equation*} B=\left( \frac{F^{\prime \prime }(\alpha _{t}^{2}o\phi .\frac{\left\vert d\phi \right\vert ^{2}}{2})}{F^{\prime }(\alpha _{t}^{2}o\phi .\frac \left\vert d\phi \right\vert ^{2}}{2})}\alpha _{t}^{2}o\phi -\frac{F^{\prime \prime }\left( \frac{\left\vert d\phi \right\vert ^{2}}{2}\right) } F^{\prime }\left( \frac{\left\vert d\phi \right\vert ^{2}}{2}\right) \right) \phi _{v}\geq 0 \end{equation* and the $F$ stress-energy tensor $S_{g}^{F}\left( \phi \right) $ of $\phi $ fulfills \begin{equation*} S_{g}^{F}\left( \gamma _{t}o\phi \right) \geq \theta \left( \alpha _{t}^{2}o\phi \right) .S_{g}^{F}\left( \phi \right) \text{ } \end{equation* where $\theta $ is a real positive function and $\gamma _{t}$ is the one parameter group of conformal transformations induced by the vector field \overline{v}$ ( defined above ) on the euclidean sphere $S^{n}$ and $\alpha _{t}$ is given by (\ref{1}). \begin{example} \end{example} The function $F(t)=\frac{1}{p}\left( 2t\right) ^{\frac{p}{2}}$ for $p=2$ and $p\geq 4$ and $t\geq 0$ is admissible. Indeed, \ for $F(t)=\frac{1}{p}\left( 2t\right) ^{\frac{p}{2}}$ we have $B=0$ and for any conformal diffeomorphism $\gamma $ on the euclidean sphere, we have \begin{equation*} S_{g}^{F}\left( \gamma o\phi \right) =\frac{1}{2}\left\vert d\left( \gamma _{t}o\phi \right) \right\vert ^{p}g-\frac{p}{2}\left\vert d\left( \gamma _{t}o\phi \right) \right\vert ^{p-2}\phi ^{\ast }can \end{equation* so if we let $\left\vert d\left( \gamma _{t}o\phi \right) \right\vert ^{2}=\alpha _{t}^{2}o\phi .\left\vert d\phi \right\vert ^{2}$, we ge \begin{equation*} S_{g}^{F}\left( \gamma _{t}o\phi \right) =\alpha _{t}^{2}o\phi .\left( \frac 1}{2}\left\vert d\phi \right\vert ^{p}g-\frac{p}{2}\left\vert d\phi \right\vert ^{p-2}\phi ^{\ast }can\right) \end{equation* \begin{equation*} =\alpha _{t}^{2}o\phi .S_{g}^{F}\left( \phi \right) \text{.} \end{equation* The $F(t)=1+at-e^{-t}$, for $t\in \left[ 0,+\infty \right[ $ where a=\max_{x\in M}\frac{\left\vert d\phi \right\vert ^{2}}{2}$ is admissible provided that the conformal diffeomorphism on the euclidean sphere $S^{n}$ is contracting that means that the function $\phi _{v}$ given in the expression of (\ref{1}) is nonnegative and the stress-energy tensor S_{g}\left( \phi \right) =\frac{\left\vert d\phi \right\vert ^{2}}{2}g-\phi ^{\ast }can$ of $\phi $ is positive. Indeed, we have \begin{equation*} B=\left( -\alpha _{t}^{2}o\phi \frac{e^{-\alpha _{t}^{2}o\phi \frac \left\vert d\phi \right\vert ^{2}}{2}}}{a+e^{-\alpha _{t}^{2}o\phi \frac \left\vert d\phi \right\vert ^{2}}{2}}}+\frac{e^{-\frac{\left\vert d\phi \right\vert ^{2}}{2}}}{a+e^{-\frac{\left\vert d\phi \right\vert ^{2}}{2}} \right) \phi _{v} \end{equation* Putting $u=\alpha _{t}^{2}o\phi \in \left] 0,1\right] $, we consider the function $\varphi \left( u\right) =-$.$u\frac{e^{-u\frac{\left\vert d\phi \right\vert ^{2}}{2}}}{a+e^{-u\frac{\left\vert d\phi \right\vert ^{2}}{2}}} \frac{e^{-\frac{\left\vert d\phi \right\vert ^{2}}{2}}}{a+e^{-\frac \left\vert d\phi \right\vert ^{2}}{2}}},$we ge \begin{equation*} \varphi ^{\prime }\left( u\right) =\left( \frac{\left\vert d\phi \right\vert ^{2}}{2}u-a-e^{-\left\vert d\phi \right\vert ^{2}u}\right) \frac{e^{-\frac \left\vert d\phi \right\vert ^{2}}{2}}}{\left( a+e^{-\frac{\left\vert d\phi \right\vert ^{2}}{2}u}\right) ^{2}} \end{equation* and it is obvious that $\varphi ^{\prime }(u)\leq 0$, hence $\varphi $ is a decreasing function on $\left] 0,1\right] $ i.e. $\varphi (u)\geq \varphi (1)=0$. Consequently $B\geq 0$. No \begin{equation*} S_{g}^{F}\left( \gamma _{t}o\phi \right) \left( X;X\right) =\left( a+e^{ \frac{\left\vert d\left( \gamma _{t}o\phi \right) \right\vert ^{2}}{2 }\right) \frac{\left\vert d\left( \gamma _{t}o\phi \right) \right\vert ^{2}} 2}g\left( X,X\right) \end{equation* \begin{equation*} -\left[ \left( a+e^{-\frac{\left\vert d\left( \gamma _{t}o\phi \right) \right\vert ^{2}}{2}}\right) -\frac{\left\vert d\left( \gamma _{t}o\phi \right) \right\vert ^{2}}{2}e^{-\frac{\left\vert d\left( \gamma _{t}o\phi \right) \right\vert ^{2}}{2}}\right] \left( \gamma _{t}o\phi \right) ^{\ast }can\left( X,X\right) . \end{equation* \begin{equation*} =\alpha _{t}^{2}o\phi \left( a+e^{-\alpha _{t}^{2}o\phi \frac{\left\vert d\phi \right\vert ^{2}}{2}}\right) \frac{\left\vert d\phi \right\vert ^{2}}{ }g\left( X,X\right) \end{equation* \begin{equation*} -\alpha _{t}^{2}o\phi \left[ \left( a+e^{-\alpha _{t}^{2}o\phi \frac \left\vert d\phi \right\vert ^{2}}{2}}\right) -\alpha _{t}^{2}o\phi \frac \left\vert d\phi \right\vert ^{2}}{2}e^{-\alpha _{t}^{2}o\phi \frac \left\vert d\phi \right\vert ^{2}}{2}}\right] \phi ^{\ast }can\left( X,X\right) \end{equation* \begin{equation*} =\alpha _{t}^{2}o\phi \left( a+e^{-\frac{\left\vert d\left( \gamma _{t}o\phi \right) \right\vert ^{2}}{2}}\right) \left[ \frac{1}{2}\left\vert d\phi \right\vert ^{2}g\left( X,X\right) -\phi ^{\ast }can\left( X,X\right) \right] \end{equation* \begin{equation*} +\alpha _{t}^{2}o\phi \frac{\left\vert d\phi \right\vert ^{2}}{2}e^{-\alpha _{t}^{2}o\phi \frac{\left\vert d\phi \right\vert ^{2}}{2}}\phi ^{\ast }can\left( X,X\right) . \end{equation* An other example is the following function $F(t)=\left( 1+2t\right) ^{\alpha }$ where $0<\alpha <1$, the $F$-energy is the $\alpha $-energy of Sacks-Uhlenbeck ( see \cite{22}). In fact \begin{equation*} B=\left( \alpha -1\right) \left( \frac{1}{1+\alpha _{t}^{2}o\phi .\left\vert d\phi \right\vert ^{2}}\alpha _{t}^{2}o\phi -\frac{1}{1+\left\vert d\phi \right\vert ^{2}}\right) \phi _{v} \end{equation* \begin{equation*} =\frac{\left( \alpha -1\right) \left( \alpha _{t}^{2}o\phi -1\right) } \left( 1+\alpha _{t}^{2}o\phi .\left\vert d\phi \right\vert ^{2}\right) \left( 1+\left\vert d\phi \right\vert ^{2}\right) }\phi _{v}\geq 0 \end{equation* provided that $\phi _{v}\geq 0$. And for vector field $X$ on $M$, we have \begin{equation*} S_{g}^{F}\left( \gamma _{t}o\phi \right) \left( X;X\right) =2\alpha \left( 1+\alpha _{t}^{2}o\phi \left\vert d\phi \right\vert ^{2}\right) ^{\alpha -1}\alpha _{t}^{2}o\phi \frac{\left\vert d\phi \right\vert ^{2}}{2}g\left( X,X\right) - \end{equation* \begin{equation*} \left[ 2\alpha \left( 1+\alpha _{t}^{2}o\phi \left\vert d\phi \right\vert ^{2}\right) ^{\alpha -1}+2\alpha \left( \alpha -1\right) \left( 1+\alpha _{t}^{2}o\phi \left\vert d\phi \right\vert ^{2}\right) ^{\alpha -2}\alpha _{t}^{2}o\phi .\left\vert d\phi \right\vert ^{2}\right] \alpha _{t}^{2}o\phi .\phi ^{\ast }can\left( X,X\right) \end{equation* \begin{equation*} =\left[ 2\alpha \left( 1+\alpha _{t}^{2}o\phi \left\vert d\phi \right\vert ^{2}\right) ^{\alpha -1}\alpha _{t}^{2}o\phi \left( \frac{\left\vert d\phi \right\vert ^{2}}{2}g\left( X,X\right) -\phi ^{\ast }can\left( X,X\right) \right) \right. + \end{equation* \begin{equation*} \left. 2\alpha \left( 1-\alpha \right) \left( 1+\alpha _{t}^{2}o\phi \left\vert d\phi \right\vert ^{2}\right) ^{\alpha -2}\alpha _{t}^{2}o\phi \left\vert d\phi \right\vert ^{2}.\phi ^{\ast }can\left( X,X\right) \right] \end{equation* and taking account of the positivity of the stress-energy tensor of S_{g}\left( \phi \right) =\frac{1}{2}\left\vert d\phi \right\vert ^{2}g-\phi ^{\ast }can$ and the fact that $\phi _{v}\geq 0$, we infer that \begin{equation*} S_{g}^{F}\left( \gamma _{t}o\phi \right) \left( X,X\right) \geq \alpha _{t}^{4}o\phi .S_{g}^{F}\phi \left( X,X\right) \text{.} \end{equation*} \begin{remark} $\phi _{v}\geq 0$ occurs for example if $\phi (M)$ is included in the positive half-sphere $S^{n+}=\left\{ x\in S^{n}:\left\langle x,v\right\rangle \geq 0\right\} $. \end{remark} In this section we state the following result \begin{theorem} \label{th2} Let $F:\left[ 0,+\infty \right[ \rightarrow \left[ 0,+\infty \right[ $ be an admissible function and $\phi $ be an $F$-harmonic map from a compact $m$-Riemannian manifold $\left( M,g\right) $ ($m\geq 2$) into the Euclidean sphere $S^{n}$ ($n\geq 2$). Suppose that the $F$ stress- energy tensor $S_{g}^{F}\left( \phi \right) $ is positive ( resp. positively defined). Then for any conformal diffeomorphism $\gamma $ on $S^{n}$, E_{F}\left( \gamma o\phi \right) \leq E_{F}\left( \phi \right) $ ( resp. E_{F}\left( \gamma o\phi \right) <E_{F}\left( \phi \right) $ ). \end{theorem} \begin{remark} In case $F(t)=\frac{1}{p}\left( 2t\right) ^{\frac{p}{2}}$, $p=2$ or $p\geq 4$ the condition $\phi _{v}\geq 0$ is not needed since $B=0$, so our result recover the ones by El-Soufi in \cite{16} and \cite{18}. \end{remark} To prove Theorem \ref{th2}, we need the following lemmas \begin{lemma} \label{lem1} Let $\phi :(M,g)\rightarrow \left( N,h\right) $ be a smooth map and $\gamma $ be a conformal diffeomorphism on $N$, then the $F$-tension of the map $\gamma o\phi $, is given b \begin{equation*} \tau _{F}\left( \gamma o\phi \right) =2\alpha ^{-1}o\phi .F^{\prime }(\alpha ^{2}o\phi \frac{\left\vert d\phi \right\vert ^{2}}{2})d\gamma \left( d\phi _{v}-\frac{\left\vert d\phi \right\vert ^{2}}{2}\nabla \alpha o\phi \right) \end{equation* \begin{equation*} +fd\gamma \left( \tau _{F}\left( \phi \right) \right) +d\gamma \left( F^{\prime }\left( \frac{\left\vert d\phi \right\vert ^{2}}{2}\right) d\phi \left( \nabla f\right) \right) \text{.} \end{equation* where $f=\frac{F^{\prime }\left( \alpha ^{2}o\phi \frac{\left\vert d\phi \right\vert ^{2}}{2}\right) }{F^{\prime }\left( \frac{\left\vert d\phi \right\vert ^{2}}{2}\right) }$ and $\gamma ^{\ast }can=\alpha ^{2}can$. \end{lemma} \begin{proof} We follow closely the proof in \cite{18 \begin{equation*} \tau _{F}\left( \gamma o\phi \right) =trace_{g}\nabla \left( F^{\prime }\left( \frac{\left\vert d\left( \gamma o\phi \right) \right\vert ^{2}}{2 \right) d\left( \gamma o\phi \right) \right) =F^{\prime }\left( \frac \left\vert d\left( \gamma o\phi \right) \right\vert ^{2}}{2}\right) trace\left( \nabla d\left( \gamma o\phi \right) \right) \end{equation* \begin{equation*} +d\left( \gamma o\phi \right) \left( \nabla F^{\prime }\left( \frac \left\vert d\left( \gamma o\phi \right) \right\vert ^{2}}{2}\right) \right) \end{equation* where $\nabla F^{\prime }\left( \frac{\left\vert d\left( \gamma o\phi \right) \right\vert ^{2}}{2}\right) $ is the gradient of $F^{\prime }\left( \frac{\left\vert d\left( \gamma o\phi \right) \right\vert ^{2}}{2}\right) $ in $M$. Since $\gamma $ is a conformal diffeomorphism on $S^{n}$, we have \begin{equation*} \tau _{F}\left( \gamma o\phi \right) =F^{\prime }\left( \alpha ^{2}o\phi \frac{\left\vert d\phi \right\vert ^{2}}{2}\right) \tau \left( \gamma o\phi \right) +d\left( \gamma o\phi \right) \left( \nabla F^{\prime }\left( \alpha ^{2}o\phi \frac{\left\vert d\phi \right\vert ^{2}}{2}\right) \right) \end{equation* \begin{equation*} =F^{\prime }\left( \alpha ^{2}o\phi \frac{\left\vert d\phi \right\vert ^{2}} 2}\right) \left( trace_{g}\nabla ^{\gamma }d\gamma \left( d\phi ,d\phi \right) +d\gamma .\tau \left( \phi \right) \right) +d\left( \gamma o\phi \right) \left( \nabla F^{\prime }\left( \alpha ^{2}o\phi \frac{\left\vert d\phi \right\vert ^{2}}{2}\right) \right) \end{equation* \begin{equation*} =F^{\prime }\left( \alpha ^{2}o\phi \frac{\left\vert d\phi \right\vert ^{2}} 2}\right) \left( trace_{g}\nabla ^{\gamma }d\gamma \left( d\phi ,d\phi \right) +\frac{1}{F^{\prime }\left( \frac{\left\vert d\phi \right\vert ^{2}} 2}\right) }d\gamma \left( \tau _{F}\left( \phi \right) \right) \right) \end{equation* \begin{equation*} -\frac{F^{\prime }(\alpha ^{2}o\phi \frac{\left\vert d\phi \right\vert ^{2}} 2})}{F^{\prime }\left( \frac{\left\vert d\phi \right\vert ^{2}}{2}\right) d\left( \gamma o\phi \right) \left( \nabla F^{\prime }\left( \frac \left\vert d\phi \right\vert ^{2}}{2}\right) \right) +d\left( \gamma o\phi \right) \left( \nabla F^{\prime }\left( \alpha ^{2}o\phi \frac{\left\vert d\phi \right\vert ^{2}}{2}\right) \right) \text{.} \end{equation* Putting $\ f=\frac{F^{\prime }(\alpha ^{2}o\phi \frac{\left\vert d\phi \right\vert ^{2}}{2})}{F^{\prime }\left( \frac{\left\vert d\phi \right\vert ^{2}}{2}\right) }$ we get \begin{equation*} \tau _{F}\left( \gamma o\phi \right) =F^{\prime }(\alpha ^{2}o\phi \frac \left\vert d\phi \right\vert ^{2}}{2})trace_{g}\nabla ^{\gamma }d\gamma \left( d\phi ,d\phi \right) +fd\gamma \left( \tau _{F}\left( \phi \right) \right) \end{equation* \begin{equation*} +F^{\prime }\left( \frac{\left\vert d\phi \right\vert ^{2}}{2}\right) d\left( \gamma o\phi \right) \left( \nabla f\right) \text{.} \end{equation*} Now since $\gamma :\left( N,\gamma ^{\ast }can\right) \rightarrow \left( N,can\right) $ is an isometry then, if $\ \widetilde{\nabla }$ denotes the connection corresponding to $\gamma ^{\ast }can$, we have \begin{equation*} \nabla ^{\gamma }d\gamma (X,Y)=d\gamma \left( \widetilde{\nabla _{X}Y-\nabla _{X}Y\right) \end{equation* and since ( see \cite{18}) \begin{equation*} \widetilde{\nabla }_{X}Y-\nabla _{X}Y=\alpha ^{-1}\left( \left\langle X,\nabla \alpha \right\rangle Y+\left\langle Y,\nabla \alpha \right\rangle X-\left\langle X,Y\right\rangle \nabla \alpha \right) \end{equation* we obtai \begin{equation*} trace_{g}\nabla ^{\gamma }d\gamma \left( d\phi ,d\phi \right) =2\alpha ^{-1}o\phi .d\gamma \left( d\phi \left( \nabla \alpha o\phi \right) -\frac \left\vert d\phi \right\vert ^{2}}{2}\nabla \alpha o\phi \right) \text{.} \end{equation*} Finally we infer tha \begin{equation*} \tau _{F}\left( \gamma o\phi \right) =2\alpha ^{-1}o\phi .F^{\prime }(\alpha ^{2}o\phi \frac{\left\vert d\phi \right\vert ^{2}}{2})d\gamma \left( d\phi \left( \nabla \alpha o\phi \right) -\frac{\left\vert d\phi \right\vert ^{2}} 2}\nabla \alpha o\phi \right) \end{equation* \begin{equation*} +fd\gamma \left( \tau _{F}\left( \phi \right) \right) +F^{\prime }\left( \frac{\left\vert d\phi \right\vert ^{2}}{2}\right) d\gamma od\phi \left( \nabla f\right) \text{.} \end{equation*} \end{proof} \begin{lemma} \label{lem2} Let $\phi $ be an $F$-harmonic map from an $m$-dimensional Riemannian manifold $\left( M,g\right) $ ($m\geq 2$) into the Euclidean unit sphere $\left( S^{n},can\right) $ ($n\geq 2$). Then for any $v\in R^{n+1}-\left\{ 0\right\} $ and any $t_{o}\in R$ we hav \begin{equation*} \frac{d}{dt}E_{F}\left( \gamma _{t}^{v}o\phi \right) _{t=t_{o}}= \end{equation* \begin{equation*} -2\frac{sht_{o}}{\left\vert v\right\vert }\int_{M}\alpha _{t_{o}}^{3}o\phi .F^{\prime }\left( \alpha _{t_{o}}^{2}o\phi .\frac{\left\vert d\phi \right\vert ^{2}}{2}\right) \left( \left\vert d\phi \right\vert ^{2}\left\vert \overline{v}o\phi \right\vert ^{2}-\left\vert d\phi _{v}\right\vert ^{2}\right) dv_{g} \end{equation* \begin{equation*} -\int_{M}\alpha _{t_{o}}^{2}o\phi .F^{\prime }\left( \frac{\left\vert d\phi \right\vert ^{2}}{2}\right) \left\langle d\phi \left( \nabla f_{t_{o}}\right) ,\overline{v}o\phi \right\rangle dv_{g} \end{equation* where $f_{t_{o}}=\frac{F^{\prime }\left( \alpha _{t_{o}}^{2}o\phi \frac \left\vert d\phi \right\vert ^{2}}{2}\right) }{F^{\prime }\left( \frac \left\vert d\phi \right\vert ^{2}}{2}\right) }$ , $\gamma ^{\ast }can=\alpha _{t_{o}}^{2}can$ and \begin{equation*} \alpha _{t_{o}}=\frac{\left\vert v\right\vert }{\phi _{v}sht_{o}+\left\vert v\right\vert cht_{o}}\text{.} \end{equation*} \end{lemma} \begin{proof} Recall that the first variation formula of the $F$-energy is given b \begin{equation*} \frac{d}{dt}E_{F}\left( \gamma _{t}^{v}o\phi \right) _{t=t_{o}}=-\int_{M}\left\langle \tau _{F}\left( \gamma _{t_{o}}^{v}o\phi \right) ,\overline{v}o\left( \gamma _{to}^{v}o\phi \right) \right\rangle dv_{g}\text{.} \end{equation* By Lemma \ref{lem1} and the fact that $\phi $ is $F$-harmonic we get \begin{equation*} \frac{d}{dt}E_{F}\left( \gamma _{t}^{v}o\phi \right) _{t=t_{o}}= \end{equation* \begin{equation*} -\int_{M}2\alpha _{t_{o}}o\phi .F^{\prime }\left( \alpha _{t_{o}}^{2}o\phi \frac{\left\vert d\phi \right\vert ^{2}}{2}\right) \left\langle \left( \nabla \alpha _{t_{o}}o\phi \right) ^{T}-\frac{\left\vert d\phi \right\vert ^{2}}{2}\nabla \alpha _{t_{o}}o\phi ,\overline{v}o\phi \right\rangle dv_{g} \end{equation* \begin{equation*} -\int_{M}F^{\prime }\left( \frac{\left\vert d\phi \right\vert ^{2}}{2 \right) \left\langle d\phi \left( \nabla f_{t_{o}}\right) ,\overline{v}o\phi \right\rangle dv_{g} \end{equation* and since ( see \cite{18} ) \begin{equation} \nabla \alpha _{t_{o}}^{v}=-\frac{\left( \alpha _{t_{o}}^{v}\right) ^{2}} \left\vert v\right\vert }sht_{o}\overline{v} \label{3} \end{equation we have \begin{equation} \left\langle \nabla \alpha _{t_{o}}o\phi ,\overline{v}o\phi \right\rangle = \frac{\left( \alpha _{t_{o}}^{v}\right) ^{2}}{\left\vert v\right\vert sht_{o}\left\vert \overline{v}o\phi \right\vert ^{2}\text{.} \label{4} \end{equation Now let $\left( e_{1},...,e_{m}\right) $ be an orthogonal basis on $M \begin{equation*} \left\langle \left( \nabla \alpha _{t_{o}}o\phi \right) ^{T},\overline{v o\phi \right\rangle =\sum_{i=1}^{m}\left\langle \nabla \alpha _{t_{o}}o\phi ,d\phi (e_{i})\right\rangle \left\langle \overline{v}o\phi ,d\phi (e_{i})\right\rangle \end{equation* \begin{equation*} =-\frac{sht_{o}}{\left\vert v\right\vert }\left( \alpha _{t_{o}}o\phi \right) ^{2}\sum_{i=1}^{m}\left\langle \overline{v}o\phi ,d\phi (e_{i})\right\rangle ^{2} \end{equation* \begin{equation*} =-\frac{sht_{o}}{\left\vert v\right\vert }\left( \alpha _{t_{o}}o\phi \right) ^{2}\left\vert d\phi _{v}\right\vert ^{2}\text{.} \end{equation*} Hence \begin{equation*} \frac{d}{dt}E_{F}\left( \gamma _{t}^{v}o\phi \right) _{t=t_{o}}= \end{equation* \begin{equation*} -2\frac{sht_{o}}{\left\vert v\right\vert }\int_{M}\alpha _{t_{o}}^{3}o\phi .F^{\prime }\left( \alpha _{t_{o}}^{2}o\phi .\frac{\left\vert d\phi \right\vert ^{2}}{2}\right) \left( \frac{\left\vert d\phi \right\vert ^{2}}{ }\left\vert \overline{v}o\phi \right\vert ^{2}-\left\vert d\phi _{v}\right\vert ^{2}\right) dv_{g} \end{equation* \begin{equation*} -\int_{M}\alpha _{t_{o}}^{2}o\phi .F^{\prime }\left( \frac{\left\vert d\phi \right\vert ^{2}}{2}\right) \left\langle d\phi \left( \nabla f_{t_{o}}\right) ,\overline{v}o\phi \right\rangle dv_{g}\text{.} \end{equation*} \end{proof} We set \begin{equation*} g(t)=\int_{M}\alpha _{t}^{2}o\phi .F^{\prime }\left( \frac{\left\vert d\phi \right\vert ^{2}}{2}\right) \left\langle d\phi \left( \nabla f_{t}\right) \overline{v}o\phi \right\rangle dv_{g}\text{.} \end{equation*} \begin{lemma} \label{lem3} \begin{equation*} g(t)= \end{equation* \begin{equation*} \int_{M}\alpha _{t}^{3}o\phi .F^{\prime \prime }(\alpha _{t}^{2}o\phi .\frac \left\vert d\phi \right\vert ^{2}}{2})\left\vert d\phi \right\vert ^{2}\left\langle d\phi \left( \nabla \left( \alpha _{t_{o}}o\phi \right) \right) ,\overline{v}o\phi \right\rangle dv_{g} \end{equation* \begin{equation*} +\int_{M}\alpha _{t}^{2}o\phi .\left( F^{\prime \prime }(\alpha _{t}^{2}o\phi .\frac{\left\vert d\phi \right\vert ^{2}}{2})\alpha _{t}^{2}o\phi -\frac{F^{\prime }(\alpha _{t}^{2}o\phi .\frac{\left\vert d\phi \right\vert ^{2}}{2})}{F^{\prime }\left( \frac{\left\vert d\phi \right\vert ^{2}}{2}\right) }F^{\prime \prime }\left( \frac{\left\vert d\phi \right\vert ^{2}}{2}\right) \right) \frac{\phi _{v}}{\left\vert v\right\vert }\left\vert d\phi \right\vert ^{2}dv_{g}\text{.} \end{equation*} \end{lemma} \begin{proof} First, we compute $\nabla f_{t} \begin{equation*} \nabla f_{t}=\frac{F^{\prime \prime }(\alpha _{t}^{2}o\phi .\frac{\left\vert d\phi \right\vert ^{2}}{2})}{F^{\prime }\left( \frac{\left\vert d\phi \right\vert ^{2}}{2}\right) }\left( \alpha _{t}o\phi .\left\vert d\phi \right\vert ^{2}\nabla \left( \alpha _{t}o\phi \right) +\alpha _{t}^{2}o\phi .\left\langle \nabla d\phi ,d\phi \right\rangle \right) \end{equation* \begin{equation*} -\frac{F^{\prime }(\alpha _{t}^{2}o\phi .\frac{\left\vert d\phi \right\vert ^{2}}{2})}{F^{\prime }\left( \frac{\left\vert d\phi \right\vert ^{2}}{2 \right) ^{2}}F^{\prime \prime }\left( \frac{\left\vert d\phi \right\vert ^{2 }{2}\right) \left\langle \nabla d\phi ,d\phi \right\rangle \end{equation* \begin{equation*} =\frac{F^{\prime \prime }(\alpha _{t}^{2}o\phi .\frac{\left\vert d\phi \right\vert ^{2}}{2})}{F^{\prime }\left( \frac{\left\vert d\phi \right\vert ^{2}}{2}\right) }\alpha _{t}o\phi \left\vert d\phi \right\vert ^{2}\nabla \left( \alpha _{t}o\phi \right) \end{equation* \begin{equation*} +\left( \frac{F^{\prime \prime }(\alpha _{t}^{2}o\phi .\frac{\left\vert d\phi \right\vert ^{2}}{2})}{F^{\prime }\left( \frac{\left\vert d\phi \right\vert ^{2}}{2}\right) }\alpha _{t}^{2}o\phi -\frac{F^{\prime }(\alpha _{t}^{2}o\phi .\frac{\left\vert d\phi \right\vert ^{2}}{2})}{F^{\prime }\left( \frac{\left\vert d\phi \right\vert ^{2}}{2}\right) ^{2}}F^{\prime \prime }\left( \frac{\left\vert d\phi \right\vert ^{2}}{2}\right) \right) \left\langle \nabla d\phi ,d\phi \right\rangle \text{.} \end{equation* Then \begin{equation*} g(t)=\int_{M}\alpha _{t_{o}}^{2}o\phi .F^{\prime }\left( \frac{\left\vert d\phi \right\vert ^{2}}{2}\right) \left\langle d\phi \left( \nabla f_{t_{o}}\right) ,\overline{v}o\phi \right\rangle dv_{g} \end{equation* \begin{equation*} =\int_{M}\alpha _{t_{o}}^{3}o\phi .F^{\prime \prime }(\alpha _{t_{o}}^{2}o\phi .\frac{\left\vert d\phi \right\vert ^{2}}{2})\left\vert d\phi \right\vert ^{2}\left\langle d\phi \left( \nabla \left( \alpha _{t_{o}}o\phi \right) \right) ,\overline{v}o\phi \right\rangle dv_{g} \end{equation* \begin{equation} +\int_{M}\alpha _{t_{o}}^{2}o\phi .\left( F^{\prime \prime }(\alpha _{t_{o}}^{2}o\phi .\frac{\left\vert d\phi \right\vert ^{2}}{2})\alpha _{t_{o}}^{2}o\phi -\frac{F^{\prime }(\alpha _{t_{o}}^{2}o\phi .\frac \left\vert d\phi \right\vert ^{2}}{2})}{F^{\prime }\left( \frac{\left\vert d\phi \right\vert ^{2}}{2}\right) }F^{\prime \prime }\left( \frac{\left\vert d\phi \right\vert ^{2}}{2}\right) \right) \label{3''} \end{equation \begin{equation*} \times \left\langle d\phi \left( \nabla \left\vert d\phi \right\vert ^{2}\right) ,\overline{v}o\phi \right\rangle dv_{g}\text{.} \end{equation* Let $\left\{ e_{1},...,e_{m}\right\} $ be a basis of $T_{x}M$ which diagonalizes $\phi ^{\ast }can$, we have \begin{equation*} \left\langle d\phi \left( \nabla \frac{\left\vert d\phi \right\vert ^{2}}{2 \right) ,\overline{v}o\phi \right\rangle =\left\langle \nabla _{e_{i}}d\phi ,d\phi \right\rangle \left\langle \overline{v}o\phi ,d\phi (e_{j})\right\rangle \left\langle d\phi (e_{i}),d\phi (e_{j})\right\rangle \end{equation* \begin{equation*} =\left\langle \nabla _{e_{i}}d\phi ,d\phi \right\rangle \left\langle \overline{v}o\phi ,d\phi (e_{j})\right\rangle \phi ^{\ast }can\left( e_{i},e_{j}\right) \end{equation* \begin{equation*} =\left\langle \nabla _{\overline{v}o\phi }d\phi \left( e_{j}\right) ,d\phi \left( e_{j}\right) \right\rangle \end{equation* \begin{equation*} =\left\langle \nabla _{d\phi \left( e_{j}\right) }\overline{v}o\phi ,d\phi \left( e_{j}\right) \right\rangle +\left\langle \left[ \overline{v}o\phi ,d\phi \left( e_{j}\right) \right] ,d\phi \left( e_{j}\right) \right\rangle \text{.} \end{equation* Likewise we ge \begin{equation*} \left\langle \left[ \overline{v}o\phi ,d\phi \left( e_{j}\right) \right] ,d\phi \left( e_{j}\right) \right\rangle =-\frac{1}{2}\frac{d}{dt}\mid _{t=0}\gamma _{t}^{\ast }\left\vert d\phi \left( e_{j}\right) \right\vert ^{2} \end{equation* \begin{equation*} =-\frac{1}{2}\frac{d}{dt}\mid _{t=0}\alpha _{t}^{2}\left\vert d\phi \left( e_{j}\right) \right\vert ^{2} \end{equation* and taking account of (\ref{1}) we obtain that \begin{equation*} \left\langle \left[ \overline{v}o\phi ,d\phi \left( e_{j}\right) \right] ,d\phi \left( e_{j}\right) \right\rangle =\frac{\phi _{v}}{\left\vert v\right\vert }\left\vert d\phi \left( e_{j}\right) \right\vert ^{2} \end{equation* so we infer that \begin{equation*} \left\langle d\phi \left( \nabla \frac{\left\vert d\phi \right\vert ^{2}}{2 \right) ,\overline{v}o\phi \right\rangle =\frac{\phi _{v}}{\left\vert v\right\vert }\left\vert d\phi \right\vert ^{2}+\left\langle \nabla _{e_{j} \overline{v}o\phi ,d\phi \left( e_{j}\right) \right\rangle \end{equation* and \begin{equation*} \left\langle \nabla _{e_{j}}\overline{v}o\phi ,d\phi \left( e_{j}\right) \right\rangle =\nabla _{e_{j}}\left\langle \overline{v}o\phi ,d\phi \left( e_{j}\right) \right\rangle -\left\langle \overline{v}o\phi ,\nabla _{e_{j}}d\phi \left( e_{j}\right) \right\rangle \end{equation* \begin{equation*} =\nabla _{e_{j}}\left\langle v,d\phi \left( e_{j}\right) \right\rangle -\left\langle \overline{v}o\phi ,\nabla _{e_{j}}d\phi \left( e_{j}\right) \right\rangle \end{equation* \begin{equation*} =\left\langle v,\nabla _{e_{j}}d\phi \left( e_{j}\right) \right\rangle -\left\langle \overline{v}o\phi ,\nabla _{e_{j}}d\phi \left( e_{j}\right) \right\rangle \end{equation* \begin{equation*} =\left\langle v-\overline{v}o\phi ,\nabla _{e_{j}}d\phi \left( e_{j}\right) \right\rangle =0\text{.} \end{equation* Henc \begin{equation} \left\langle d\phi \left( \nabla \frac{\left\vert d\phi \right\vert ^{2}}{2 \right) ,\overline{v}o\phi \right\rangle =\frac{\phi _{v}}{\left\vert v\right\vert }\left\vert d\phi \right\vert ^{2}\text{.} \label{4'} \end{equation} \end{proof} Now set \begin{equation*} \varphi (t_{o})=2\frac{sht_{o}}{\left\vert v\right\vert }\int_{M}\alpha _{t_{o}}^{3}o\phi .F^{\prime }\left( \alpha _{t_{o}}^{2}o\phi .\frac \left\vert d\phi \right\vert ^{2}}{2}\right) \left( -\frac{\left\vert d\phi \right\vert ^{2}}{2}\left\vert \overline{v}o\phi \right\vert ^{2}+\left\vert d\phi _{v}\right\vert ^{2}\right) dv_{g} \end{equation* \begin{equation*} -\int_{M}\alpha _{t_{o}}^{3}o\phi .F^{\prime \prime }(\alpha _{t_{o}}^{2}o\phi .\frac{\left\vert d\phi \right\vert ^{2}}{2})\frac \left\vert d\phi \right\vert ^{2}}{2}\left\langle d\phi \left( \nabla \left( \alpha _{t_{o}}o\phi \right) \right) ,\overline{v}o\phi \right\rangle dv_{g} \end{equation* and since by (\ref{3}) we have $_{{}} \begin{equation*} \left\langle d\phi \left( \nabla \left( \alpha _{t_{o}}o\phi \right) \right) ,\overline{v}o\phi \right\rangle =-\frac{sht_{o}}{\left\vert v\right\vert \alpha _{t_{o}}^{2}o\phi \left\vert d\phi _{v}\right\vert ^{2} \end{equation* we ge \begin{equation} \varphi (t_{o})=2\frac{sht_{o}}{\left\vert v\right\vert }\int_{M}\alpha _{t_{o}}^{3}o\phi .\left[ \left( F^{\prime }\left( \alpha _{t_{o}}^{2}o\phi \frac{\left\vert d\phi \right\vert ^{2}}{2}\right) +\alpha _{t_{o}}^{2}o\phi .\frac{\left\vert d\phi \right\vert ^{2}}{2}F^{\prime \prime }\left( \alpha _{t_{o}}^{2}o\phi .\frac{\left\vert d\phi \right\vert ^{2}}{2}\right) \right) \left\vert d\phi _{v}\right\vert ^{2}\right. \label{5} \end{equation \begin{equation*} \left. -F^{\prime }\left( \alpha _{t_{o}}^{2}o\phi .\frac{\left\vert d\phi \right\vert ^{2}}{2}\right) \frac{\left\vert d\phi \right\vert ^{2}}{2 \left\vert \overline{v}o\phi \right\vert ^{2}\right] dv_{g}\text{.} \end{equation* or \begin{equation*} \varphi (t_{o})=-2\frac{sht_{o}}{\left\vert v\right\vert }\int_{M}\alpha _{t_{o}}^{3}o\phi .S_{g}^{F}\left( \gamma _{t}^{v}o\phi \right) dv_{g}\text{ } \end{equation*} \begin{proof} ( of Theorem \ref{th2}) \ Recall ( see \cite{13} ) that for any conformal diffeomorphism $\gamma $ of the unit sphere $S^{n}$ there exist an isometry r\in O\left( n+1\right) $, a real number $t\geq 0$ and a vector $v\in R^{n+1}-\left\{ 0\right\} $ such that $\gamma =ro\gamma _{t}^{v}$, so it suffices to consider $\gamma _{t}^{v}$ with $t\geq 0$ and $v\in R^{n+1}-\left\{ 0\right\} $. On the other hand \begin{equation*} \frac{d}{dt}E_{F}\left( \gamma _{t}^{v}o\phi \right) =\varphi (t)+\chi (t) \end{equation* wher \begin{equation*} \chi (t)=-\int_{M}\alpha _{t}^{2}o\phi .\left( F^{\prime \prime }(\alpha _{t}^{2}o\phi .\frac{\left\vert d\phi \right\vert ^{2}}{2})\alpha _{t}^{2}o\phi -\frac{F^{\prime }(\alpha _{t}^{2}o\phi .\frac{\left\vert d\phi \right\vert ^{2}}{2})}{F^{\prime }\left( \frac{\left\vert d\phi \right\vert ^{2}}{2}\right) }F^{\prime \prime }\left( \frac{\left\vert d\phi \right\vert ^{2}}{2}\right) \right) \frac{\phi _{v}}{\left\vert v\right\vert }\left\vert d\phi \right\vert ^{2}dv_{g} \end{equation* and $\varphi \left( t\right) $ is given by (\ref{5}). Now, since the function $F$ is admissible we infer that$\ \chi (t)\leq 0$ . Since the $F$ energy-stress tensor $S_{g}^{F}\left( \phi \right) $of $\phi $ is positive ( resp. positive defined ) by assumption and \begin{equation*} \geq S_{g}^{F}\left( \phi \right) \end{equation* so the tensor field $S_{g}^{F}\left( \gamma _{t}o\phi \right) $ is positive ( resp. positive defined ). Consequently $\varphi (t)\leq 0$ ( resp. \varphi (t)<0$ ) for any $\ t\geq 0$ and the proof of Theorem \ref{th2} is complete. \end{proof}
\section{Introduction} A moving contact line occurs at the location where two ostensibly immiscible fluids and a solid meet. It arises in a wide range of both natural and technological processes, from insects walking on water~\cite{insect} and the wetting properties of plant leaves~\cite{leaves} to coating~\cite{coating}, inkjet printing~\cite{inkjet,inkjet2} and oil recovery~\cite{oil}. In addition to its crucial role in wide-ranging applications, it remains a persistent problem, a long-standing and fundamental challenge in the field of fluid dynamics, despite its apparent simplicity at first sight (see \textit{e.g.}~review articles~\cite{Dussan79,deGennesrev,blake2006physics,BonnEggers}). Not surprisingly it has been investigated extensively, both experimentally and theoretically, for several decades. At the heart of the moving contact line problem is that, when treated classically as two immiscible fluids moving along a solid surface satisfying the no-slip condition, there is no solution due to the multivalued velocity at the contact line,~\cite{DussanDavis,ShikhSingualarities06}. This is known most famously in the literature as the problem of a non-integrable stress singularity, a result published a few decades ago along with the nonphysical prediction that an infinite force is required to submerge a solid object~\cite{HuhScriv71}. The resolution of the problem may have initially appeared straightforward. The no-slip condition at the wall could not be satisfied, thus some form of slip in the contact line vicinity should be allowed. Navier-slip, written down in the early 19th Century~\cite{Navier}, was a prime candidate, a form of which was suggested in the concluding remarks of~\cite{HuhScriv71}. The fact that wetting and the moving contact line remain an open debate and a fruitful research area is largely due to the particular microscale ingredients that may alleviate the problem being numerous and hotly debated---see \textit{e.g.}~the wide range of discussion articles recently,~\cite{epj_entire}. Various alternative models to slip at the wall include: a precursor film ahead of the contact line~\cite{SchwartzEley}; rheological effects~\cite{WeidnerSchwartz}; treating surfaces as separate thermodynamic entities with dynamic surface tensions~\cite{Shikh93} (see a recent critical investigation of this model,~\cite{My_Shikh}); introducing numerical slip~\cite{RenardySlip}; including evaporative fluxes~\cite{ColinetRednikov}; and considering the interface to be diffuse, numerical work reviewed \textit{e.g.}~in~\cite{anderson_rev}. Here, we examine analytically a diffuse-interface model without any other ingredients, being both self-consistent and physically relevant: rather than considering a sharp fluid-fluid interface as a surface of zero thickness where quantities (\textit{e.g.}~density) are, in general, discontinuous, it considers the interface to have a non-zero, finite, thickness with quantities varying smoothly but rapidly, in agreement with developments from the statistical mechanics of liquids community (\textit{e.g.}~\cite{EvansReview,yatsyshin1}). The fluid density $\bar\rho$ then acts as an order parameter such that in the sharp-interface limit the two bulk fluids satisfy $\bar\rho=\rho_L$ and $\bar\rho=\rho_V=0$, being liquid and vapour respectively, where we consider the behaviour of the system with vapour phase of negligible density, noting that an equivalent double-well free energy form (to be described later, in sect. \ref{sec:ps}), with zero bulk vapour density, is also used by Pismen and Pomeau~\cite{PismenPomeau} and is physically relevant for liquid-gas systems. The interface between liquid and gas may then be defined as the location where $\bar\rho=(\rho_L+\rho_V)/2$. Diffuse-interface models have been popular for numerical implementation as tracking of the fluid-fluid interface is not required in the resulting free-boundary problem, instead the interface is inferred from density field contours. For solid-liquid-gas systems the seminal study of Seppecher~\cite{seppecher} is often referred to when suggesting that diffuse-interface models resolve the moving contact line problem. Whilst Seppecher's work contains some discussion of the asymptotics, the analysis was largely incomplete, with asymptotic regions being probed without careful justification and the crucial behaviour close to the contact line only investigated numerically (a number of constraints were also imposed, \textit{e.g.}~90$^\circ$ contact angles and fluids of equal viscosity). Full numerical simulations for the liquid-gas problem have also been undertaken (\textit{e.g.}~in~\cite{BriantYeomansEarly,BriantYeomans1}); binary fluids have also been examined using diffuse-interface methods of a different form, where a coupled Cahn--Hilliard equation models the diffusion between the two components~\cite{KhatavkarJFM,DingSpeltJFM,YueZhouFeng}. Here, we undertake an analytical investigation by considering the contact line behaviour for a liquid-gas system with two basic elements: (a) the interface has a finite thickness, which is expected from statistical mechanics studies as noted earlier, and (b) the no-slip condition is applied at the wall. This diffuse-interface model then resolves the moving contact line problem without the need to model any further physical effects from the microscale. An important ramification of this analysis is that the wetting boundary condition used in conjunction with diffuse-interface in existing numerical studies of liquid-gas systems needs to be appropriately modified, otherwise it leads to a density gradient orthogonal to the wall at large distances from the contact line. The possibility of density variations such as these are often included when disjoining pressure models are considered, where many studies utilise the long-wave (or lubrication) approximation, \textit{e.g.}~\cite{deGennesrev,oron1997long}. In the 1985 review of de Gennes \cite{deGennesrev}, the thickness of precursor films was discussed. For complete wetting on a dry substrate, nanometric films, decaying ahead of the contact line, were predicted. A recent study of intermolecular forces in the contact line region using approaches from statistical mechanics, namely density-functional theory~\cite{antoniojfm}, demonstrated that for the case of partially wetting fluids, a constant-thickness nanometric film of a few molecular diameters is adsorbed in front of the contact line. More recent experiments have also been performed (\textit{e.g.}~to probe the dynamics of such nanoscale films \cite{KavPre11}). In the macro/mesoscopic setting of this work, and to isolate the contribution of the diffuse-interface model without depletion/enrichment near the solid substrate, we thus predominately consider the case where the bulk densities are valid up to the walls (but other cases are considered), and we make no assumptions on thin films---rather considering arbitrary finite contact angles. \begin{figure}[t] \centering \includegraphics{Fig1.eps}\ \caption{Sketch of the diffuse-interface model near a wall.} \label{fig:diffinterface} \end{figure} \section{Problem specification} \label{sec:ps} Consider a fluid in the upper half $(\bar x,\bar y)$-plane~$\Omega$ with a solid surface~$\partial\Omega$ at $\bar y=0$; see Fig.~\ref{fig:diffinterface}. The free energy of the system has contributions from an isothermal fluid with a Helmholtz free energy functional and from a wall energy $f_w=f_w(\bar\rho)$, thus given by $\mathscr{F} = \int_\Omega \bar{\mathcal{L}} \, \textrm{d}\Omega + \int_{\partial\Omega} f_w \, \textrm{d} A$, where \begin{equation} \bar{\mathcal{L}} = \bar\rho \bar{f}(\bar\rho) + {K}|\bm{\nabla}\bar\rho|^2/{2} -\bar{G}\bar\rho, \label{eq:origL} \end{equation} with $\bar{G}$ the chemical potential, $K$ a gradient energy coefficient (assumed constant for simplicity), and $\bar\rho \bar f(\bar\rho)$ a double well potential chosen to give the two equilibrium states $\bar\rho=\{0,\rho_L\}$. Such a form for the free energy and associated diffuse-interface approximations has been used by numerous authors for wetting problems such as Seppecher \cite{seppecher}, Pismen and Pomeau \cite{PismenPomeau}, and Pismen \cite{Pismenmeso}; see also the review by de~Gennes \cite{deGennesrev}. The effect of the non-local terms neglected in the local approximation to obtain such a free energy has been considered at equilibrium in the aforementioned studies \cite{Pismenmeso,antoniojfm}, where the long-range intermolecular interactions are responsible for an algebraic decay of the density profile away from the interface instead of the exponential one as predicted here (seen later, in \refe{eq:cak0}). In our dynamic situation, we focus on the local approximation to elucidate the contact line behaviour in a simplified, yet widely used setting. The density field augments the usual hydrodynamic equations via the capillary (or Korteweg) stress tensor $\bar{\tens{T}}$ through \begin{equation} \bar f(\bar\rho) = \frac{K}{2\epsilon^2}\bar\rho\lrsq{ 1-\frac{\bar\rho}{\rho_L} }^2, \quad \bar{\tens{T}} = \bar{\mathcal{L}}\tens{I} - \bm{\nabla} \bar\rho \otimes \pfrac{\bar{\mathcal{L}}}{(\bm{\nabla} \bar\rho)},\label{eq:origf} \end{equation} where $\tens{I}$ is the identity tensor, and with $\epsilon$ being the interface width, and $\bar{\tens{T}}$ arising from Noether's theorem,~\cite{anderson_rev}. Following \cite{seppecher}, \cite{anderson_rev} and \cite{PismenPomeau}, using \refe{eq:origL} and coupling in the compressible Stokes equations (assuming creeping flow) the capillary tensor with the usual viscous stress tensor $\bar\bm{\tau}$, taken as deviatoric for simplicity, yields \begin{gather} \pilfrac{\bar\rho}{\bar{t}} + \bm{\nabla}\bm{\cdot}(\bar\rho\bar\vect{u}) = 0, \qquad \bm{\nabla} \bm{\cdot} (\bar{\tens{T}} + \bar\bm{\tau}) =0, \nonumber \\ \bar{\tens{T}} = \left( \bar\rho \bar f(\bar\rho) + {K}|\bm{\nabla}\bar\rho|^2/2 - \bar{G}\bar\rho \right)\tens{I} - K\bm{\nabla}\bar\rho\otimes\bm{\nabla}\bar\rho, \nonumber \\ \bar\bm{\tau} = \bar\mu(\bar\rho)\left[(\bm{\nabla}\bar\vect{u})+(\bm{\nabla}\bar\vect{u})^\mathrm{T} - {2}(\bm{\nabla}\bm{\cdot}\bar\vect{u})\tens{I}/3\right], \nonumber\\ \bar{G} = -K\nabla^2\bar\rho + \pilfrac{}{\bar\rho}(\bar\rho \bar f(\bar\rho)),\label{eq:gesdim} \end{gather} where $\bar\vect{u}$ and $\bar\mu(\bar\rho)$ are the fluid velocity and viscosity, respectively, $\bar{t}$ is time, and the thermodynamic pressure is given by $\bar p=\bar\rho^2 \bar f'(\bar\rho)$. The form for $\bar{G}$ arises from the Euler-Lagrange equation corresponding to the free energy, and we take $\bar\mu(\bar\rho) = \mu_L \bar\rho/\rho_L$, giving the viscosities for the two equilibrium states as $\mu_L$ and $\mu_V = 0$. On the solid surface $\partial\Omega$, with normal $\vect{n}_w$, we impose \begin{equation} \bar{\vect{u}} = \bar{\vect{u}}_w, \qquad K \vect{n}_w \bm{\cdot} \bm{\nabla} \bar\rho + \bar f'_w(\bar\rho) = 0,\label{eq:origbcs} \end{equation} with wall velocity $\bar{\vect{u}}_w=(-V,0)$ in Cartesian coordinates. The first condition is classical no-slip, whilst the second arises by variational arguments and is termed the \textit{natural} (or \textit{wetting}) boundary condition, \cite{YueZhouFeng04}. $\bar f_w(\bar\rho)$ is chosen to satisfy Young's law at the contact line, with solid-liquid, solid-gas and liquid-gas surface tensions $\bar f_w(\rho_L)=\sigma_L$, $\bar f_w(\rho_V)=\sigma_V$ and $\sigma$ respectively, and with contact angle $\theta_S$. A cubic is the lowest-order polynomial required for the wall free energy to be minimised by the bulk densities, to prevent depletion/enrichment away from the contact line, \textit{i.e.}~$\bar f'_w(\bar\rho_L)=\bar f'_w(0)=0$. Whilst cubic forms are used for binary fluid problems, \textit{e.g.}~\cite{jacqmin,YueZhouFeng}, this is unlike the linear forms proposed in previous studies for liquid-gas problems \cite{seppecher,BriantYeomansEarly,BriantYeomans1,qianliqgas}, and allows us to consider a diffuse-interface model without further physical effects from the microscale (although this is relaxed in the following section). We define \begin{equation*} \bar f_w(\bar\rho) = \lrsq{{\rho_L^{-3}\sigma\cos\theta_S} ( 4\bar\rho^3 -6\bar\rho^2\rho_L + \rho_L^3 ) + \sigma_V+\sigma_L }/2, \end{equation*} giving \begin{equation*} \bar f'_w(\bar\rho) = -{6\sigma}\bar\rho\left( \rho_L - \bar\rho \right)\cos\theta_S/{\rho_L^3}, \end{equation*} and $\bar f_w(0)-\bar f_w(\rho_L) = \sigma\cos\theta_S$, with Young's law thus satisfied. It is noteworthy that \refe{eq:origbcs}(b) may be replaced by a constant density condition if a precursor film/disjoining pressure model is to be considered, \textit{i.e.}~$\bar{\rho}=\rho_a$ on $\partial\Omega$ (as used in \cite{PismenPomeau}, and considered in the following section). Finally, for a one-dimensional density profile $\bar\rho(z)$ in equilibrium (\textit{i.e.}~with $\bar{G}=0$ required by our choice of $\bar\rho\bar{f}(\bar\rho)$ to have equally stable bulk fluids, and $\bar{\vect{u}}=0$), the surface tension across the interface is \begin{equation} \sigma = K \int_{-\infty}^\infty\left( \frac{\textrm{d}\bar{\rho}}{\textrm{d}\bar{z}} \right)^2\, \textrm{d}\bar{z}=\frac{K\bar\rho_L^2}{6\epsilon}\label{eq:dimsurfaceT}, \end{equation} see \textit{e.g.}~\cite{CH58}, and we note Eqs. \refe{eq:gesdim}--\refe{eq:origbcs}, without the specific choices for $\bar{\mu}(\bar\rho)$, $\bar{f}(\bar\rho)$ and $\bar{f}_w(\bar\rho)$ are as derived/used in \cite{seppecher,anderson_rev,PismenPomeau}. To nondimensionalise, we use typical length, velocity and density scales $X$, $V$ and $\rho_L$ respectively. The pressure and viscous stress are scaled with $\ilfrac{\mu_L V}{X}$, and the capillary stress with $\ilfrac{K\rho_L^2}{(\epsilon X)}$. Finally, $\bar{G}$ is scaled with $\ilfrac{K\rho_L}{(\epsilon X)}$ and $\bar{f}$ with $\ilfrac{K\rho_L}{\epsilon^2}$. The governing equations then contain the nondimensional parameters \begin{equation*} \operatorname{Cn} = \ilfrac{\epsilon}{X}, \quad\mbox{and}\quad \operatorname{Ca}_k = \ilfrac{\mu_L V \epsilon}{(K \rho_L^2)}, \end{equation*} being the Cahn number and a modified Capillary number based on the model parameter $K$, respectively. Nondimensional variables are denoted as their dimensional counterparts with bars dropped, \textit{e.g.}~nondimensional bulk densities are $\rho=\{0,1\}$. $\operatorname{Ca}_k$ is related to the usual Capillary number, $\operatorname{Ca}=\mu_L V / \sigma = 6\operatorname{Ca}_k$, through \refe{eq:dimsurfaceT}. The governing equations \refe{eq:gesdim} in nondimensional form are \begin{gather} \pilfrac{\rho}{t} + \bm{\nabla}\bm{\cdot}(\rho\vect{u}) = 0, \quad \tens{M} = {\operatorname{Ca}_k}^{-1}\tens{T} + \bm{\tau}, \quad \bm{\nabla} \bm{\cdot} \tens{M} = 0, \nonumber\\ \tens{T} = \left( {\operatorname{Cn}}^{-1} \rho f(\rho) + {\operatorname{Cn}}|\bm{\nabla}\rho|^2/2 - G\rho \right)\tens{I} -\operatorname{Cn}\bm{\nabla}\rho\otimes\bm{\nabla}\rho, \nonumber\\ \bm{\tau} = \rho\lrsq{(\bm{\nabla}\vect{u})+(\bm{\nabla}\vect{u})^\mathrm{T} - {2}(\bm{\nabla}\bm{\cdot}\vect{u})\tens{I}/{3}}, \nonumber\\ G = -\operatorname{Cn}\nabla^2\rho + {\operatorname{Cn}}^{-1}\pilfrac{}{\rho}(\rho f(\rho)),\label{eq:nd1} \end{gather} where $\tens{M}$ is introduced as the total stress tensor, $p=(\operatorname{Cn}\operatorname{Ca}_k)^{-1}\rho^2f'(\rho)$, and from \refe{eq:origf}: \begin{align} f(\rho) = \frac{\rho}{2}(1-\rho)^2, \quad \mbox{giving} \quad p = \frac{\rho^2(1-3\rho)(1-\rho)}{2\operatorname{Cn}\operatorname{Ca}_k}.\label{eq:nd2} \end{align} On the solid surface $\partial\Omega$, from \refe{eq:origbcs}, we have \begin{align} \vect{u} = \vect{u}_w, \qquad \operatorname{Cn} \vect{n}_w \bm{\cdot} \bm{\nabla} \rho = \cos\theta_S(1-\rho)\rho ,\label{eq:ndge2} \end{align} where $\vect{u}_w=(-1,0)$, and $\vect{n}_w=(0,-1)$, in Cartesian components. We can rewrite the governing equations \refe{eq:nd1} for steady flows as \begin{align} \bm{\nabla}\bm{\cdot}(\rho\vect{u}) &= 0, \nonumber\\ \bm{\nabla} p &= (2\operatorname{Cn}\operatorname{Ca}_k)^{-1}\bm{\nabla} [\rho^2(1-3\rho)(1-\rho)],\nonumber\\ 0&=\operatorname{Ca}_k^{-1} {\operatorname{Cn}} \ \rho\bm{\nabla}(\nabla^2\rho) - \bm{\nabla} p \nonumber\\ &\qquad + \rho[\bm{\nabla}^2\vect{u} + \bm{\nabla}(\bm{\nabla}\bm{\cdot}\vect{u})/3] \nonumber\\ & \qquad + [(\bm{\nabla}\vect{u})+(\bm{\nabla}\vect{u})^\mathrm{T} - 2(\bm{\nabla}\bm{\cdot}\vect{u})\tens{I}/3] \bm{\nabla}\rho ,\label{eq:ndge1} \end{align} and the boundary conditions on $\partial\Omega$ remain as above, in \refe{eq:ndge2}. We initially consider the equilibrium behaviour of the system, corresponding to $\operatorname{Ca}_k\ll1$, to provide a basis for comparison when the dynamic behaviour is analysed, where eqs.~\refe{eq:ndge2}--\refe{eq:ndge1} are thus reduced to \begin{equation*} 2{\operatorname{Cn}^2}\rho\bm{\nabla}(\nabla^2\rho) = \bm{\nabla} \lrsq{\rho^{2}(1-3\rho)(1-\rho)}, \end{equation*} subject to $-\operatorname{Cn}\pilfrac{\rho}{y} = \cos\theta_S(1-\rho)\rho$ at $y=0$, and with $\rho\to\{0,1\}$ and $\nabla^2\rho\to0$ as $x\to\pm\infty$ to obtain the expected bulk behaviour. The solution subject to the above conditions is \begin{equation} \rho = \lr{1-\tanh\lrsq{(2\operatorname{Cn})^{-1}(x\sin\theta_S+y\cos\theta_S) }}/2, \label{eq:cak0} \end{equation} having also fixed the interface at $\rho=1/2$. This profile is planar and at angle $\theta_S$, shown in Fig.~\ref{fig:eqd} in inner variables (where $\{x,y\} = \operatorname{Cn} \{\td{x},\td{y}\}$, for comparison to forthcoming plots). \begin{figure}[t] \centering \includegraphics{Fig2.eps}\ \caption{The equilibrium density behaviour (contours) near the contact line for $\theta_S=\pi/4$, from eq.~\refe{eq:cak0} in inner variables.} \label{fig:eqd} \end{figure} For physical systems, the scale over which the density varies between liquid and gas is much smaller than the macroscopic length scale, and hence $\operatorname{Cn}\ll1$. The asymptotic behaviour as $\operatorname{Cn}\to0$ is known as the sharp-interface limit, and understanding of it is of vital importance when considering diffuse-interface models as classical continuum models should be recovered if correct predictions are to be found. A careful asymptotic analysis of the outer solution away from the interface, and of the interfacial region away from the wall (using body fitted coordinates), shows that the expected sharp-interface equations (the Stokes equations with no-slip and the usual capillary surface stress conditions) are indeed recovered. Consider now the inner region near to both interface and wall in polar coordinates with $r=O(\operatorname{Cn})$. The scaling $\bm{\nabla} = \operatorname{Cn}^{-1}\inin{\bm{\nabla}}$ (\textit{i.e.}~$r=\operatorname{Cn}\inin{r}$, with inner variables denoted with tildes) retains all terms in the governing equations and boundary conditions \refe{eq:ndge2}--\refe{eq:ndge1}, giving a complete dominant balance. The steady governing equations, from \refe{eq:ndge1}, in this inner region are thus \begin{align} \inin\bm{\nabla}\bm{\cdot}(\inin\rho\inin\vect{u}) = 0,\label{eq:ndinge1} \end{align} and \begin{align} &\operatorname{Ca}_k^{-1} \inin{\rho}\inin\bm{\nabla}(\inin\nabla^2\inin\rho) - (2\operatorname{Ca}_k)^{-1}\inin\bm{\nabla} [\inin\rho^2(1-3\inin\rho)(1-\inin\rho)] \nonumber\\ & \qquad\qquad + \inin\rho[\inin\bm{\nabla}^2\inin\vect{u} + \inin\bm{\nabla}(\inin\bm{\nabla}\bm{\cdot}\inin\vect{u})/3] \nonumber\\ & \qquad\qquad + [(\inin\bm{\nabla}\inin\vect{u})+(\inin\bm{\nabla}\inin\vect{u})^\mathrm{T} - 2(\inin\bm{\nabla}\bm{\cdot}\inin\vect{u})\tens{I}/3] \inin\bm{\nabla}\inin\rho = 0,\label{eq:ndinge2} \end{align} and on the solid surface $\partial\Omega$, the boundary conditions are \begin{align} \inin\vect{u} = \inin\vect{u}_w, \qquad \vect{n}_w \bm{\cdot} \inin\bm{\nabla} \inin\rho = \cos\theta_S(1-\inin\rho)\inin\rho , \end{align} where in polar coordinates, and with $\inin{u}$ and $\inin{v}$ being the radial and angular velocity components, these reduce to \begin{align} \inin{u} &= -1, & \inin{v} &= 0, & -\frac{1}{\inin{r}}\pfrac{\inin{\rho}}{\theta} &= \cos\theta_S(1-\inin{\rho})\inin{\rho},\label{eq:ndinge3} \intertext{on $\theta=0$, and} \inin{u} &= 1, & \inin{v} &= 0, & \frac{1}{\inin{r}}\pfrac{\inin{\rho}}{\theta} &= \cos\theta_S(1-\inin{\rho})\inin{\rho},\label{eq:ndinge4} \end{align} on $\theta=\pi$. Of particular interest is the behaviour as the contact line is approached---the location where a stress singularity or no solution (due to multivalued velocity) arises in the classical formulation of the problem. To consider the asymptotic solution as the contact line is approached, we expand \begin{equation*} \{\inin\rho,\inin{u},\inin{v}\} = \sum_{i=0}^\infty \{ \inin\rho_i(\theta), \inin{u}_i(\theta), \inin{v}_i(\theta) \} \, \inin{r}^i, \end{equation*} in \refe{eq:ndinge1}--\refe{eq:ndinge2}, and find at leading order \begin{equation*} {\inin{\rho}_0}(\inin{u}_0 + \inin{v}_0') = - {\inin{\rho}_0'}\inin{v}_0, \qquad \inin{\rho}_0''' = 0, \qquad \inin{\rho}_0'' = 0, \end{equation*} subject to $\inin{\rho}_0'=0$ on $\theta=\{0,\pi\}$, from \refe{eq:ndinge3}--\refe{eq:ndinge4}. The density is solved as $\inin{\rho}_0=1/2$, having imposed its expected value at the interface. To find the leading-order velocities, we continue to first order in the governing equations \refe{eq:ndinge1}--\refe{eq:ndinge2}, where \begin{gather*} \inin{u}_0 = -\inin{v}_0', \qquad \operatorname{Ca}_k(\inin{v}_0''' + \inin{v}_0') + \inin{\rho}_1'' + \inin{\rho}_1 = 0, \nonumber\\ \operatorname{Ca}_k(\inin{v}_0'' + \inin{v}_0) - \inin{\rho}_1''' - \inin{\rho}_1' = 0. \end{gather*} with the wetting and no-slip conditions from \refe{eq:ndinge3}--\refe{eq:ndinge4} being \begin{align*} \inin{u}_0 &= -1, & \inin{v}_0 &= 0, & -\inin{\rho}_1' &= \cos(\theta_S)/4, \intertext{on $\theta=0$, and} \inin{u}_0 &= 1, & \inin{v}_0 &= 0, & \inin{\rho}_1' &= \cos(\theta_S)/4, \end{align*} on $\theta=\pi$. We also assert that the profile must be planar at these very small distances to the contact line for a well-defined microscopic contact angle in the Young equation---requiring $\inin{\rho}(\inin{r},\pi-\theta_S)=\ilfrac{1}{2}$, at least up to this first-order correction, and leading to \begin{equation*} \inin{\rho}_1 = -\frac{\sin\theta_S\cos\theta+\cos\theta_S\sin\theta}{4}, \ \lr{\begin{array}{c} \inin{u}_0 \\ \inin{v}_0 \end{array}} = \lr{\begin{array}{c} -\cos\theta \\ \sin\theta \end{array}}. \end{equation*} We now consider the stresses and pressure, which in inner variables are scaled with $\operatorname{Cn}^{-1}$ (readily seen from eqs.~\refe{eq:nd1}--\refe{eq:nd2}), as their singular behaviour in the classical model of \cite{HuhScriv71} is the hallmark of the moving contact line problem. The total stress components in polar coordinates and in inner variables are \begin{align*} \inin{\tens M}_{{\inin{r}}{\inin{r}}} &= {\operatorname{Ca}_k}^{-1}[ \{{\inin{\rho}^2(1-\inin{\rho})^2} + \lr{\pilfrac{\inin{\rho}}{\theta}}^2/{\inin{r}^2} - \lr{\pilfrac{\inin{\rho}}{\inin{r}}}^2\} /2 \nonumber\\ & + \inin{\rho}\lrcur{\ppilfrac{\inin{\rho}}{\inin{r}} + \ilfrac{\pilfrac{\inin{\rho}}{\inin{r}}}{\inin{r}} + \ilfrac{\ppilfrac{\inin{\rho}}{\theta}}{\inin{r}^2} - {\inin{\rho}(1-\inin{\rho})(1-2\inin{\rho})}} ] \nonumber\\ & + \inin{\rho}\lrsq{{4}\pilfrac{\inin{u}}{\inin{r}}/{3}-{2}\lr{ \inin{u} + \pilfrac{\inin{v}}{\theta}}/{(3 \inin{r})}}, \\ \inin{\tens M}_{\inin{r}\theta} &= -{\operatorname{Ca}_k}^{-1}\pilfrac{\inin{\rho}}{\inin{r}} \ \pilfrac{\inin{\rho}}{\theta}/\inin{r} + \inin{\rho}\lrsq{\lr{\pilfrac{\inin{u}}{\theta} - \inin{v}}/\inin{r} + \pilfrac{\inin{v}}{\inin{r}}}, \\ \inin{\tens M}_{\theta\theta} &= {\operatorname{Ca}_k}^{-1}[ \{{\inin{\rho}^2(1-\inin{\rho})^2} - \lr{\pilfrac{\inin{\rho}}{\theta}}^2/\inin{r}^2 + \lr{\pilfrac{\inin{\rho}}{\inin{r}}}^2\} /2 \nonumber\\ & + \inin{\rho}\lrcur{ \ppilfrac{\inin{\rho}}{\inin{r}} + \pilfrac{\inin{\rho}}{\inin{r}}/\inin{r} + \ppilfrac{\inin{\rho}}{\theta}/{\inin{r}^2} - {\inin{\rho}(1-\inin{\rho})(1-2\inin{\rho})}} ] \nonumber\\& + \inin{\rho}\lrsq{- {2}\pilfrac{ \inin{u}}{ \inin{r}}/3 + {4}\lr{ \inin{u} + \pilfrac{ \inin{v}}{\theta}}/{(3 \inin{r})}}, \end{align*} so by substituting in our results, we find at leading order $\inin{\tens{M}} = O(1)$, as all $O(1/\inin{r})$ terms cancel. To obtain the precise form of the stresses, the second-order terms in the governing equations are needed. The pressure in this inner region is given by $\inin p = \inin\rho^2(1-3\inin\rho)(1-\inin\rho)/(2\operatorname{Ca}_k)$, so that as $\inin r \to 0$ it satisfies $\inin p = -1/(32\operatorname{Ca}_k) + O(\inin r)$, being finite at the contact line. Continuing with these second-order terms, we find the density and velocity corrections \begin{align*} \inin\rho_2 &= {C_{\inin \rho_1}+C_{\inin \rho_2}\cos(2\theta)}, \\ \lr{\begin{array}{c} \inin{u}_1 \\ \inin{v}_1 \end{array}} &= \lr{\begin{array}{c} {\sin\theta_S(\cos(2\theta)-1)/4 - C_{\inin v}\sin(2\theta)} \\ {-\sin\theta_S\sin(2\theta)/4 + C_{\inin v}(1-\cos(2\theta))} \end{array}} , \end{align*} arising through solving the governing equations \begin{gather*} \inin{u}_1 = -\sin(\theta_S)/4 - \inin{v}_1'/2, \quad \inin{v}_1' = -\inin{v}_1'''/4, \\ \inin{\rho}_2' = -\inin{\rho}_2'''/4 , \end{gather*} and boundary conditions \begin{align*} \inin{u}_1 = \inin{v}_1 = \inin{\rho}_2' = 0 \qquad \mbox{on }\theta=\{0,\pi\}, \end{align*} at this order, and where the arbitrary constants $C_{\inin \rho_1}$, $C_{\inin \rho_2}$ and $C_{\inin v}$ would be set by the full solution of the inner problem. A possible flow scenario where $C_{\inin \rho_1}=-0.1$, $C_{\inin \rho_2}=0.3$, $C_{\inin v}=-1$, for $\theta_S=\pi/4$ is shown in Fig.~\ref{fig:fo_plots}(a). All of these results allow us to determine the leading-order stress components as \begin{align*} \inin{\tens M}_{\inin{r}\inin{r}} &= (M_1\cos(2\theta) - M_2\sin(2\theta) + M_3)/32 + O(\inin r), \\ \inin{\tens M}_{\inin{r}\inin{\theta}} &= -(M_2\cos(2\theta) + M_1\sin(2\theta))/32 + O(\inin r), \\ \inin{\tens M}_{\inin{\theta}\inin{\theta}} &= (M_2\sin(2\theta) -M_1\cos(2\theta) + M_3)/32 + O(\inin r), \end{align*} where \begin{align*} M_1 &= {\cos(2\theta_S)}/{\operatorname{Ca}_k}+8\sin\theta_S, \\ M_2 &= 32C_{\inin v}+{\sin(2\theta_S)}/{\operatorname{Ca}_k}, \\ M_3 &= {(1+64C_{\inin \rho_1})}/{\operatorname{Ca}_k}-\ilfrac{8\sin\theta_S}{3}, \end{align*} showing that the stresses are nonsingular as $\inin r \to 0$. \section{Extensions} Having demonstrated the ability of our diffuse-interface model to alleviate the moving contact line problem with no-slip applied, we now consider a number of other features of contact line flow. These extensions are to demonstrate the range of boundary conditions derived and applied in the literature, being relevant to various physical situations, and to draw comparisons between diffuse-interface models and one of the more complex continuum theories proposed to deal with contact line flows, the interface formation model of Shikhmurzaev \cite{ShikhBook}. A recent paper, \cite{My_Shikh}, critically examined this interface formation model of Shikhmurzaev \cite{ShikhBook}. There, it was shown that the model degenerates to the same macroscopic flow as slip models but it was also seen to have features that most of the simpler models such as Navier-slip do not. In particular, the interface formation model captures finite pressure behaviour, it generalises both the fluid-fluid and the fluid-solid interfaces from their classical models of being sharp and with no-slip respectively through the modelling of surface layers, and the microscopic contact angle is able to vary dynamically from its static value (with its value determined as part of the solution rather than prescribed empirically as it is sometimes for slip models, \textit{e.g.}~\cite{Hocking90} ). Finally, the fluid is able to `roll', as in a moving frame of reference there is no stagnation point at the contact line, allowing particles to reach and transfer through the contact line in finite time. Whilst these features occur at lengthscales too small to probe with current experimental ability (see discussions in \cite{My_Shikh}), it is of interest that the diffuse-interface model is capable of similar predictions, with various features added. The model studied thus far already predicts finite pressure at the contact line, and alleviates the stagnation point predicted by slip models through mass transfer. It relaxes the sharp fluid-fluid interface assumption, but has classical no-slip at the wall in contrast to the effective slip of the interface formation model. Although not necessary, this may also be relaxed through carefully prescribing a generalised Navier boundary condition (GNBC), suggesting that the slip velocity is proportional to the total tangential stress (the sum of the viscous and uncompensated Young stress---arising from the deviation of the fluid-fluid interface from its static shape), and derivable using variational arguments from the principle of minimum energy dissipation \cite{QianWangShengGNBCfirst,QianWangShengJFM}. Our diffuse-interface model also prescribes the microscopic dynamic contact angle $\theta_d$ to be equal to the static value $\theta_S$ through the wetting boundary condition. An alternative is for this condition to hold at equilibrium, with the density relaxing to it in finite time when out of equilibrium, as initially discussed (but not implemented) for binary fluids \cite{jacqmin}, and more recently used in numerical simulations \cite{QianWangShengGNBCfirst,QianWangShengJFM,YueFengEPJ,YueFengPoF}. \begin{figure}[t] \centering \includegraphics{Fig3.eps}\ \caption{The asymptotic behaviour of density (contour plots) and velocity (arrows) as the contact line is approached. The driving force in the system is the moving wall.} \label{fig:fo_plots} \end{figure} For our liquid-gas configuration, the GNBC (of Qian, Wang and Sheng \cite{QianWangShengGNBCfirst,QianWangShengJFM}) and generalised wetting boundary condition (of the variety of authors mentioned above \cite{jacqmin,QianWangShengGNBCfirst,QianWangShengJFM,YueFengEPJ,YueFengPoF}) may be considered analogously. In dimensional form the wetting boundary condition is generalised to \begin{equation} \alpha\lr{\pilfrac{\bar\rho}{\bar t} + \bar\vect{u}\bm{\cdot}\bm{\nabla}\bar\rho} = -\bar L(\bar\rho),\label{eq:wallrelaxwet} \end{equation} where $ \bar L(\bar\rho) = K \vect{n}_w \bm{\cdot} \bm{\nabla} \bar\rho + \bar f'_w(\bar\rho)$, is the wall chemical potential, and $\alpha=0$ representing instantaneous relaxation to equilibrium. A representative effect of \refe{eq:wallrelaxwet} is shown in Fig.~\ref{fig:fo_plots}(d), in comparison to \refe{eq:origbcs} in Fig.~\ref{fig:fo_plots}(a). The GNBC for this application with inverse slip length $\bar\beta$ is then \begin{equation} \bar L(\bar\rho) (\vect{t}_w\bm{\cdot}\bm{\nabla}\bar\rho) - \bar\bm{\tau}_{nt} = \bar\beta(\bar\vect{u}-\bar\vect{u}_w)\bm{\cdot}\vect{t}_w, \label{eq:dimGNBC} \end{equation} with $\vect{t}_w$ the tangent to the wall, and $\bar\bm{\tau}_{nt}$ the viscous shear stress. Note $\alpha=0$ reduces to the popular Navier-slip condition. A representative effect of the GNBC from \refe{eq:dimGNBC} is shown in Fig.~\ref{fig:fo_plots}(e), and in combination with \refe{eq:wallrelaxwet} in Fig.~\ref{fig:fo_plots}(f). In nondimensional form in polar coordinates for steady flow in the inner region, \refe{eq:wallrelaxwet} reduces to \begin{equation} \inin L(\inin\rho) = \operatorname{Ca}_k \Pi \lr{\inin u \, \pilfrac{\inin\rho}{\inin r} + {\inin v} \, \pilfrac{\inin\rho}{\theta}/{\inin r}},\label{eq:wallrelaxwetin} \end{equation} where $\inin L(\inin\rho) = \pm\pilfrac{\inin\rho}{\theta} /\inin r + \inin\rho(1-\inin\rho)\cos\theta_S$ on $\theta=\{0,\pi\}$, and with another nondimensional parameter arising, $\Pi={\alpha\rho_L^2}/({\mu_L\epsilon})$, describing the extent of the wall relaxation ($\Pi=0$ being instantaneous). Similarly \refe{eq:dimGNBC} reduces to \begin{equation} \mp \inin L(\inin\rho)\pilfrac{\inin\rho}{\inin r}/\operatorname{Ca}_k - \inin\rho\lrsq{(\pilfrac{\inin u}{\theta} - {\inin v})/{\inin r} + \pilfrac{\inin v}{\inin r}} = \beta (1\pm \inin u),\label{eq:dimGNBCin} \end{equation} where $\beta = {\bar\beta \epsilon}/{\mu_L}$ is the nondimensional slip parameter, and along with $\Pi$ are both chosen to be formed with the interface thickness $\epsilon$ such that they are considered as $O(1)$ in the limit $\operatorname{Cn}\to0$. To consider how \refe{eq:wallrelaxwet} allows for microscopic contact angle variation dependent on flow conditions, we note that \refe{eq:wallrelaxwet} at equilibrium (denoted with subscript $e$) gives \begin{equation*} \bar L(\bar\rho) = \lreval{K \vect{n}_w \bm{\cdot} \bm{\nabla} \bar\rho}_e + \bar f'_w(\rho_L/2) = 0. \end{equation*} Based on calculations in \cite{QianWangShengJFM,YueFengPoF} for the binary fluid case, we consider a steady, dynamic situation (denoted with subscript $d$), and see that \refe{eq:wallrelaxwet} implies \begin{equation*} \alpha \lreval{\bar\vect{u}\bm{\cdot}\bm{\nabla}\bar\rho}_{d} = - \lreval{K \vect{n}_w \bm{\cdot} \bm{\nabla} \bar\rho}_{d} - \bar f'_w(\rho_L/2), \end{equation*} thus using the equilibrium result and considering this at the contact line with wall velocity $\bar\vect{u}_w=-V\vect{t}_w$, we find $ V\alpha\sin\theta_d = - K (\cos\theta_d - \cos\theta_S)$, or in nondimensional form \begin{equation*} \operatorname{Ca}_k\Pi = \ilfrac{(\cos\theta_S- \cos\theta_d)}{\sin\theta_d} \approx \theta_d-\theta_S, \end{equation*} where the final approximation holds for $\theta_d -\theta_S \ll 1$, in agreement with the binary fluid case \cite{YueFengPoF}. Another consideration is the behaviour near the contact line if density gradients near the wall are permitted far from the contact line. Our wall free energy was specifically chosen to prevent this, but we may also consider the two other situations used previously in the literature, namely (i) specifying a density at the wall $\bar\rho=\rho_a$ on $\partial\Omega$, as in \cite{PismenPomeau}, and a representative effect of this shown in Fig.~\ref{fig:fo_plots}(c) and (ii) choosing a linear form in the density for the wall free energy $\bar f_w(\bar\rho) = a\bar\rho$, as in \cite{seppecher,BriantYeomansEarly,BriantYeomans1,qianliqgas}, and shown in Fig.~\ref{fig:fo_plots}(b). To consider the contact line behaviour when (i) replaces the wetting boundary condition is straightforward, but for (ii), we must understand how to impose the microscopic contact angle to compare to our previous condition. Following \cite{BriantYeomansEarly}, we use Young's law $\cos\theta_S = (\sigma_{V}-\sigma_{L})/\sigma$ and compute $\sigma_{V}$ and $\sigma_{L}$ by integrating the free energy per unit area along the corresponding interface. This gives $\cos\theta_S = [(1-A)^{3/2}-(1+A)^{3/2}]/2$, where $A = 4a\epsilon/(K\rho_L)$ is nondimensional. This may then be inverted to give the appropriate value of $A$ for a given contact angle $\theta_S$, and corresponds to the nondimensional boundary condition $ \operatorname{Cn} \vect{n}_w \bm{\cdot} \bm{\nabla} \rho = -A/4 $. Adding these features into the diffuse-interface model do not dramatically alter the contact line behaviour, but subtle differences in the asymptotic results are demonstrated, as mentioned, in Fig.~\ref{fig:fo_plots} for selected arbitrary constants, where $\operatorname{Ca}_k=0.1$ and $\theta_S=\pi/4$, and may be compared to the equilibrium situation in Fig.~\ref{fig:eqd}. The cases considered are (a) the original model \refe{eq:origbcs} without slip or wall relaxation, (b) using the linear form for $\bar{f}_w(\bar\rho)$ in the wetting boundary condition, (c) adding a precursor film at the wall (where the wetting boundary condition is replaced by $\rho=\rho_a/\rho_L=0.53$), (d) allowing finite wall relaxation, \refe{eq:wallrelaxwetin} with $\Pi=5$, (e) including the GNBC (\refe{eq:dimGNBCin} with $\beta=2$) but with $\Pi=0$, and (f) using \refe{eq:wallrelaxwetin} with $\Pi=5$ and \refe{eq:dimGNBCin} with $\beta=2$. All models behave as expected, resolving the stress and pressure singularities, and including the effects they intend, \textit{e.g.}~capturing the film in (c), increased microscopic contact angle in (d) and (f), and reduced wall velocity in (e) and (f). There are only small differences between cubic and linear wall energy forms near the contact line, mainly that the linear form shows a broader band of density variation. This hints at the important difference that will occur near the wall but far away from the interface, where density gradients will remain present for this linear form. \section{Conclusions} We have shown analytically that a diffuse-interface model is able to resolve the moving contact line problem through relaxing the interface from being sharp to thin, without need to model any further physical effects from the microscale at the contact line. Whilst slip, precursor films and finite-time wall relaxation have been considered, they are \emph{not} necessary to resolve the moving contact line problem. We believe that the present study will motivate further analytical and numerical work with diffuse-interface models, such as to consider heterogeneous walls \cite{Savva09,Savva10,RajChemHet,christophe_marc}. Of particular interest would also be the inclusion of non-local terms into the governing equations; this was considered in \cite{antoniojfm} for equilibrium wetting using a density-functional theory. \section*{Acknowledgements} We acknowledge financial support from ERC Advanced Grant No. 247031, and Imperial College through a DTG International Studentship. \bibliographystyle{abbrv}
\section*{Introduction} The theory of mixed type equations is one the principal parts of the general theory of partial differential equations. The interest for these kinds of equations arises intensively because of both theoretical and practical of their applications. Many mathematical models of applied problems require investigations of this type of equations. The actuality of the consideration of mixed type equations has been mentioned, for the first time, by S. A. Chaplygin in 1902 in his famous work ``On gas streams'' \cite{chap}. The first fundamental results in this direction was obtained in 1920-1930 by F. Tricomi \cite{tric} and S. Gellerstedt \cite{gell}. The works of M. A. Lavrent'ev \cite{lavbit}, A. V. Bitsadze \cite{bit}, \cite{bit2}, F. I. Frankl \cite{fra}, M. Protter \cite{pro} and C. Morawetz \cite{mor}, have had a great impact in this theory, where outstanding theoretical results were obtained and pointed out important practical values of them. Bibliography of the main fundamental results on this direction can be found, among others, in the monographs of A. V. Bitsadze \cite{bit2}, Y. M. Berezansky \cite{bere}, L. Bers \cite{bers}, M. S. Salakhitdinov and A. K. Urinov \cite{sal} and A. M. Nakhushev \cite{nak}. In most of the works devoted to the study of mixed type equations, the object of study was mixed elliptic-hyperbolic type equations. Comparatively, few results have been obtained on the study of mixed parabolic-hyperbolic type equations. However, this last type of equations have also numerous applications in the real life processes (see \cite{ost} for an interesting example in mechanics). The reader can found a nice example given, for the first time, by Gelfand in \cite{gel}, and connected with the movement of the gas in a channel surrounded by a porous environment. Inside the channel the movement of gas was described by the wave equation and outside by the diffusion one. Mathematic models of this kind of problems arise in the study of electromagnetic fields, in heterogeneous environment, consisting of dielectric and conductive environment for modeling the movement of a little compressible fluid in a channel surrounded by a porous medium \cite{lei}. Here the wave equation describes the hydrodynamic pressure of the fluid in the channel, and the equation of filtration-pressure fluid in a porous medium. Similar problems arise in the study of the magnetic intensity of the electromagnetic field \cite{lei}. In the last few years, the investigations on local boundary value problems, for mixed equations in domains with non-characteristic boundary data, were intensively increased. We point out that the studies made on boundary value problems for equations of mixed type, in domains with deviation from the characteristics (with non-characteristic boundary), have been originated with the fundamental works of Bitsadze \cite{bit}, where the generalized Tricomi problem (Problem M) for an equation of mixed type is discussed. In the works \cite{salb} and \cite{sal}, the analog to the Tricomi problem for a modeled parabolic-hyperbolic equation, was investigated in a domain with non-characteristic boundary in a hyperbolic part. Moreover, the uniqueness of solution and the Volterra property of the formulated problem was proved. We also refer to the recent works devoted to the study of parabolic-hyperbolic equations \cite{asho}-\cite{khub}. In the last years, the interest for considering boundary value problems of parabolic-hyperbolic type, with integral gluing condition on the line of type changing, is increasing \cite{berk}, \cite{kapm}. In the present work we study the analog to the generalized Tricomi problem with integral gluing condition on the line of type changing. We prove that the formulated problem has the Volterra property. The obtained result generalizes some previous ones from M. A. Sadybekov and G. D. Tajzhanova given in \cite{sadt}. \section*{Formulation of the problem } Let $\Omega \subset {{{\mathbb R}}^2}$ be a domain, bounded at $y > 0$ by segments $A{A_0},{A_0}{B_0},B{B_0}$ of straight lines $x = 0,y = 1,x = 1$ respectively, and at $y < 0$ by a monotone smooth curve $AC:\,y = - \gamma (x),\,0 < x < l,\, 1/2 < l < 1,\,\gamma (0) = 0,\,l + \gamma (l) = 1$ and by the segment $BC:\,x - y = 1,\,l \leq x < 1$, which is the characteristic curve of the equation $$ Lu = f(x,y),\eqno (1) $$ where $$ L\, u = \left\{ \begin{array}{l} {u_x} - {u_{yy}},\,\,\,\,y > 0, \hfill \\ \\ {u_{xx}} - {u_{yy}},\,\,\,\,y < 0. \hfill \\ \end{array} \right.\eqno (2) $$ Now we state the problem that we will consider along the paper: \textbf{Problem B.} To find a solution of the Eq.(1), satisfying boundary conditions $$ {\left. {u(x,\,\,y)} \right|_{A{A_0} \cup {A_0}{B_0}}} = 0,\eqno (3) $$ $$ {\left. {({u_x} - {u_y})} \right|_{AC}} = 0\eqno (4) $$ and gluing conditions $$ {u_x}(x, + 0) = {u_x}(x, - 0),\,\,\,\,\,{u_y}(x, + 0) = \alpha \,{u_y}(x,\, - 0) + \beta \int\limits_0^x {{u_y}(t,\, - 0)Q(x,t)\,dt} ,\,\,\,0 < x < 1,\eqno (5) $$ where $Q$ is a given function such that $Q \in {C^1}\left( {\left[ {0,1} \right] \times \left[ {0,1} \right]} \right)$, and $\alpha ,\,\,\beta \in {{\mathbb R}}$ satisfy ${\alpha ^2} + {\beta ^2} > 0$. When the curve $AC$ coincides with the characteristic one $x + y = 0$, $\alpha = 1$ and $\beta = 0$, the problem B is just the Tricomi problem for parabolic-hyperbolic equation with non-characteristic line of type changing, which has been studied in \cite{ele}. Regular solvability of the problem B with continuous gluing conditions ($\alpha = 1,\,\,\,\,\beta = 0$) have been proved, for the first time, in \cite{ber}, and strong solvability of this problem was proved in the work \cite{sadt}. Several properties, including the Volterra property of boundary problems for mixed parabolic-hyperbolic equations, have been studied in the works \cite{ber2}-\cite{berck}. We denote the parabolic part of the mixed domain $\Omega$ as $\Omega_0$ and the hyperbolic part by $\Omega_1$. A regular solution of the problem B in the domain $\Omega$ will be a function $$ u \in C(\bar \Omega ) \cap {C^1}({\Omega _0} \cup AB)\cap {C^1}({\Omega _1} \cup AC \cup AB) \cap {C^{1,2}}({\Omega _0})\, \cap {C^{2,2}}({\Omega _1}), $$ that satisfies Eq.(1) in the domains $\Omega_0$ and $\Omega_1$, the boundary conditions (3)-(4), and the gluing condition (5). Regarding the curve $AC$, we assume that $x + \gamma (x)$ is monotonically increasing. Then, rewriting it by using the characteristic variables $\xi = x + y$ and $\eta = x - y$, we have that the equation of the curve $AC$ can be expressed as $\xi = \lambda (\eta ),\,\,\,0 \leq \eta \leq 1$. \section*{Main result} \textbf{Theorem 1.} Let $\gamma \in {C^1}[0,l]$ and $Q \in {C^1}\left( {\left[ {0,1} \right] \times \left[ {0,1} \right]} \right)$. Then for any function $f \in {C^1}(\bar \Omega )$, there exists a unique regular solution of the Problem B. \verb"Proof:" By a regular solution of the problem B in the domain ${\Omega _1}$ we look for a function that fulfills the following expression $$ u(\xi ,\,\,\eta ) = \frac{1}{2}\left[ {\tau (\xi ) + \tau (\eta ) - \int\limits_\xi ^\eta {{\nu _1}(t)dt} } \right] - \int\limits_\xi ^\eta {d{\xi _1}\int\limits_{{\xi _1}}^\eta {{f_1}({\xi _1},\,{\eta _1})} d{\eta _1}},\eqno (6) $$ where $$ \xi = x + y,\,\,\,\,\eta = x - y,\,\,\,{f_1}(\xi ,\eta ) = \frac{1}{4}f(\frac{{\xi + \eta }}{2},\,\frac{{\xi - \eta }}{2}\,),\,\,\,\,\,\tau (x) = u(x,\, - 0),\,\,\,\,\,\,\,\,{\nu _1}(x) = {u_y}(x,\, - 0).\eqno (7) $$ Based on (4) from (6), using the expressions on (7), we deduce that $$ {\nu _1}(\eta ) = \tau '(\eta ) - 2\int\limits_{\lambda (\eta )}^\eta {{f_1}({\xi _1},\,\eta )} d{\xi _1},\,\,\,0 \leq \eta \leq 1.\eqno (8) $$ By virtue of the unique solvability of the first boundary problem for the heat equation (1) satisfying condition (3), and the fact that $u\left( {x,0} \right) = \tau \left( x \right)$, its solution can be represented as $$ u(x,y) = \int\limits_0^x {d{x_1}} \int\limits_0^1 {G(x - {x_1},\,y,{y_1})f({x_1},\,{y_1})} d{y_1} + \int\limits_0^x {{G_{{y_1}}}} (x - {x_1},y,0)\tau ({x_1})d{x_1},\eqno (9) $$ where $\tau (0) = 0$ and $G\left( {x,y,{y_1}} \right)$ is the Green's function related to the first boundary problem, for the heat equation in a rectangle $A{A_0}{B_0}B$, which has the form \cite{tik} $$ G(x,y,{y_1}) = \frac{1}{{2\sqrt {\pi x} }}\sum\limits_{n = - \infty }^{ + \infty } {\left[ {\exp \left\{ { - \frac{{{{(y - {y_1} + 2n)}^2}}}{{4x}}} \right\} - \exp \left\{ { - \frac{{{{(y + {y_1} + 2n)}^2}}}{{4x}}} \right\}} \right]}.\eqno (10) $$ Calculating the derivative $\frac{{\partial u}}{{\partial y}}$ in (9) and passing to the limit at $y \to 0$ we get $$ {u_y}(x, + 0) = - \int\limits_0^x {k(x - t){u_x}(t, + 0)dt} + {F_0}(x), $$ where $$ k(x) = \frac{1}{{\sqrt {\pi x} }}\sum\limits_{n = - \infty }^{ + \infty } {{e^{ - \frac{{{n^2}}}{x}}}} = \frac{1}{{\sqrt \pi {x^{\frac{1}{2}}}}} + \widetilde k(x),\eqno (11) $$ and $$ {F_0}(x) = \int\limits_0^x {d{x_1}} \int\limits_0^1 {{{\left. {{G_y}(x - {x_1},y,{y_1})} \right|}_{y = 0}}f({x_1},{y_1})d{y_1}}.\eqno (12) $$ Thus, the main functional relation between $\tau '(x)$ and ${\nu _0}(x) = {u_y}(x, + 0)$, reduced to the segment $AB$ from the parabolic part of the domain, imply that $$ {\nu _0}(x) = - \int\limits_0^x {k(x - t)\tau '(t)dt} + {F_0}(x).\eqno (13) $$ Suppose, in a first moment, that $\alpha \ne 0$. From (8) and (13), considering the gluing condition (5), we obtain the following integral equation regarding the function $\tau '(x)$: $$ \tau '(x) + \int\limits_0^x {{k_1}(x,t)} \tau '(t)dt = {F_1}(x).\eqno (14) $$ Here $$ {k_1}(x,t) = \frac{1}{\alpha }\left[ {k(x - t) + \beta Q(x,t)} \right],\eqno (15) $$ and $$ {F_1}(x) = \frac{1}{\alpha }{F_0}(x) + 2\int\limits_{\lambda (x)}^x {{f_1}({\xi _1},x)d{\xi _1}} + \frac{{2\beta }}{\alpha }\int\limits_0^x {Q(x,t)dt\int\limits_{\lambda (t)}^t {f({\xi _1},t)d{\xi _1}} }. \eqno (16) $$ Hence, the problem B is equivalent, in the sense of unique solvability, to the second kind Volterra integral equation (14). The restrictions imposed on the functions $\gamma$, $Q$, and the right hand of the Eq.(1) guarantee that, by virtue of (11) and (15), the kernel ${k_1}(x,t)$ is a kernel with weak singularity. So, we have that Eq.(14) has a unique solution and $\tau' \in {C^1}(0,1)$. Since $\tau(0)=0$, we deduce the uniqueness of the function $\tau$. Eq.(8) gives us the uniqueness of function $\nu_1$ and, as consequence, we deduce, from Eq.(6), the uniqueness of solution of problem B when $\alpha \ne 0$. Consider now the other case, i.e. $\alpha = 0$ and $\beta \ne 0$. From functional relations (8) and (13), and taking gluing condition (5) into account at $\alpha = 0$, we have $$ - \int\limits_0^x {k(x - t)\tau '} (t)dt + {F_0}(x) = \beta \int\limits_0^x {[\tau '(t) - 2\int\limits_{\lambda (t)}^t {{f_1}({\xi _1},t)d{\xi _1}} ]Q(x,t)dt} $$ or, which is the same, $$ \int\limits_0^x {\tau '(t)[k(x - t) + \beta Q(x,t)]dt} = {F_0}(x) + 2\beta \int\limits_0^x {dt\int\limits_{\lambda (t)}^t {Q(x,t){f_1}({\xi _1},t)d{\xi _1}} .} $$ Considering the representation of $k(x - t)$, the previous equation can be rewritten as follows $$ \int\limits_0^x {\frac{\tau '(t)dt}{{\left(x - t\right)}^{1/2}}} = \sqrt \pi \left[ F_0(x) + 2\beta \int\limits_0^x {dt} \int\limits_{\lambda (t)}^t {Q(x,t){f_1}(\xi _1,t)d{\xi _1}} - \int\limits_0^x {\tau'(t)\left(\widetilde{k}(x - t)+ \beta Q(x,t)\right) dt} \right].\eqno (17) $$ Since Eq.(17) is the Abel's equation, it can be solved and so we arrive at the following identity: \begin{eqnarray*} \tau '(x) &=& \displaystyle \frac{{{F_0}(0)}}{{\sqrt {\pi x} }} + \frac{1}{{\sqrt \pi }}\left\{ {\int\limits_0^x {\frac{{{F'_0}(t)dt}}{{\sqrt {x - t} }} + 2\beta \int\limits_0^x {\frac{{dt}}{{\sqrt {x - t} }}} } \frac{\partial }{{\partial t}}\int\limits_0^t {dz\int\limits_{\lambda (z)}^z {Q(t,z){f_1}({\xi _1},z)d{\xi _1}} } } \right. \hfill \\ && - \left. {\int\limits_0^x {\frac{{dt}}{{\sqrt {x - t} }}} \frac{\partial }{{\partial t}}\int\limits_0^t {\tau (z)\left[ {\widetilde k(x - t) + \beta Q(t,z)} \right]dz} } \right\}. \hfill \\ \end{eqnarray*} Considering ${F_0}(0) = 0$, after some simplifications we get $$ \tau '(x) + \int\limits_0^x {{K_0}(x,z)} \tau '(z)dz = {F_2}(x),\eqno (18) $$ where $$ {K_0}(x,z) = \frac{1}{{\sqrt \pi }}\left\{ {\frac{{Q(z,z)}}{{\sqrt {x - z} }} + \int\limits_0^{x - z} {{{(x - t)}^{ - \frac{1}{2}}}} \frac{\partial }{{\partial t}}\left[ {\widetilde k(t - z) + \beta Q(t,z)} \right]dt} \right\}, $$ \begin{eqnarray*} \hspace{3cm} {F_2}(x) &=&\frac{1}{{\sqrt \pi }}\int\limits_0^x {d{x_1}} \int\limits_0^1 {\left[ {\int\limits_0^{x - {x_1}} {\frac{{{G_{yt}}(t,{y_1},0)}}{{\sqrt {x - {x_1} - t} }}dt} } \right]f({x_1},{y_1})d{y_1}} \hfill \\ & & \hspace{11cm} (19)\\ &&+ \frac{{2\beta }}{{\sqrt \pi }}\int\limits_0^x {d{\eta _1}} \int\limits_{\lambda ({\eta _1})}^{{\eta _1}} {\left[ {\frac{{Q({\eta _1},{\eta _1})}}{{\sqrt {x - {\eta _1}} }} + \int\limits_0^{x - {\eta _1}} {\frac{{{Q_t}(t,{\eta _1})}}{{\sqrt {x - {\eta _1} - t} }}dt} } \right]} {f_1}({\xi _1},{\eta _1})d{\xi _1}. \hfill \\ \end{eqnarray*} Since the kernel ${K_0}(x,z)$ has a weak singularity, then Eq.(18) has a unique solution, and it can be represented as $$ \tau '(x) = {F_2}(x) + \int\limits_0^x {R(x,z)} {F_2}(z)dz,\eqno (20) $$ where $R(x,z)$ is the resolvent kernel of (18). As consequence, arguing as in the case $\alpha\neq 0$, we deduce, from Eq.(6), the uniqueness of solution of Problem B for $\alpha=0$ and $\beta\neq 0$, and the result is proved. \hbox to 0pt{}\hfill$\rlap{$\sqcap$}\sqcup$ In the sequel, we will deduce the exact expression of the integral kernel related to the unique solution of Problem B. To this end, we suppose, at the beginning, that $\alpha\neq 0$. Note that the unique solution of Eq.(14) can be represented as $$ \tau '(x) = \int\limits_0^x {\Gamma (x,t)} {F_1}(t)dt + {F_1}(x),\eqno (21) $$ where $\Gamma (x,t)$ is the resolvent kernel of the Eq.(14), and it is given by the recurrence formula: $$ \Gamma (x,t) = \sum\limits_{n = 1}^\infty {{{( - 1)}^n}{k_{1n}}(x,t),\,\,\,\,{k_{11}}(x,t)} = {k_1}(x,t),\,\,\,{k_{1n + 1}}(x,t) = \int\limits_0^x {{k_1}(x,z){k_{1n}}(t,z)} dz. $$ From (21), taking $\tau (0) = 0$ into account, we have that $$ \tau (x) = \int\limits_0^x {{\Gamma _1}(x,t)} {F_1}(t)dt, $$ where ${\Gamma _1}(x,t) = 1 + \int\limits_t^x {\Gamma (z,t)} dz$. From the formula (6), and considering (8), one can easily deduce that $$ u(\xi,\eta) = \tau (\xi ) + \int\limits_\xi ^\eta {d{\eta _1}} \int\limits_{\lambda ({\eta _1})}^{\eta_1} {f({\xi _1},{\eta _1})d{\xi _1}} . \eqno (22) $$ Substituting the representation of $\tau (x)$ into (22) and considering (12) and (16), after some evaluations we get \begin{eqnarray*} \hspace{2cm} u(x,y) &=& \frac{1}{\alpha }\int\limits_0^\xi {d{x_1}} \int\limits_0^1 {{G_1}(\xi - {x_1},{y_1})f({x_1},{y_1})d{y_1}} + 2\int\limits_0^\xi {d{\eta _1}} \int\limits_{\lambda ({\eta _1})}^{{\eta _1}} {{\Gamma _1}(\xi ,{\eta _1}){f_1}({\xi _1},{\eta _1})d{\xi _1}} \hfill \\ && \hspace{12cm} (23)\\ && + \frac{{2\beta }}{\alpha }\int\limits_0^\xi {d{\eta _1}} \int\limits_{\lambda ({\eta _1})}^{{\eta _1}} {{G_0}(\xi - {\eta _1},{\eta _1}){f_1}({\xi _1},{\eta _1})d{\xi _1}} + \int\limits_\xi ^\eta {d{\eta _1}} \int\limits_{\lambda ({\eta _1})}^{\eta_1} {f({\xi _1},{\eta _1})d{\xi _1},} \end{eqnarray*} where $$ {G_1}(x,{y_1}) = \int\limits_0^x {{\Gamma _1}(x,t){G_y}(t,{y_1},0)dt}, $$ and $$ {G_0}(x,\eta ) = \int\limits_0^x {Q(z + \eta ,\eta ){\Gamma _1}(x,z)dz}. $$ In an analogous way, substituting the representation of $\tau (x)$ into (9), we have \begin{eqnarray*} \hspace{2cm} u(x,y) &=& \int\limits_0^x {d{x_1}} \int\limits_0^1 {{G_2}(x - {x_1},y,{y_1})f({x_1},{y_1})d{y_1}} + 2\int\limits_0^x {d{\eta _1}} \int\limits_{\lambda ({\eta _1})}^{{\eta _1}} {{G_1}(x - {\eta _1},y){f_1}({\xi _1},{\eta _1})d{\xi _1}} \hfill \\ && \hspace{12cm} (24)\\ && + \frac{{2\beta }}{\alpha }\int\limits_0^x {d{\eta _1}} \int\limits_{\lambda ({\eta _1})}^{{\eta _1}} {{G_{01}}(x - {\eta _1},{\eta _1}){f_1}({\xi _1},{\eta _1})d{\xi _1}} , \hfill \\ \end{eqnarray*} where $$ {G_2}(x,y,{y_1}) = G(x,y,{y_1}) + \frac{1}{\alpha }\int\limits_0^x {{G_1}(t,{y_1}){G_y}(x - t,y,0)dt} $$ and $$ {G_{01}}(x,{\eta _1}) = \int\limits_0^x {{G_y}({x_1},{y_1},0){G_0}(x - {x_1},{\eta _1})d{x_1}} . $$ From (23) and (24), we arrive at the following expression $$ u(x,y) = \iint\limits_\Omega {{K_{\alpha \beta }}(x,y,{x_1},{y_1})f({x_1},{y_1})}d{x_1}d{y_1}, $$ where \begin{eqnarray*} {K_{\alpha \beta }}(x,y,{x_1},{y_1}) &=& \theta (y)\left\{ {\frac{{}}{{}}\theta ({y_1})\theta (x - {x_1}){G_2}(x - {x_1},y,{y_1}) + \theta ( - {y_1})\theta (x - {\eta _1})\left[ {\frac{{}}{{}}{G_1}(x - {\eta _1},y) } \right.} \right. \hfill \\ && \left. +{\left. {\frac{{2\beta }}{\alpha }{G_{01}}(x - {\eta _1},{\eta _1})} \right]} \right\} + \theta ( - y)\left\{ {\theta ({y_1})\theta (\xi - {x_1}){G_1}(\xi - {x_1},{y_1})} \right. \hfill \\ && +\left. \theta ( - {y_1})\left[ {\frac{1}{2}\theta (\eta - {\eta _1})\theta ({\eta _1} - \xi )\theta (\xi - {\xi _1}) } \right. {+\left. {\theta (\xi - {\eta _1})\left[ {{\Gamma _1}(\xi ,{\eta _1}) + \frac{\beta }{\alpha }{G_0}(\xi - {\eta _1},{\eta _1})} \right]} \right]} \right\}. \hfill \end{eqnarray*} Here $$ \theta (y) = \left\{ \begin{array}{l} 1,\,\,\,\,y > 0, \hfill \\ \\ 0,\,\,\,\,y < 0. \hfill \\ \end{array} \right. $$ When $\alpha=0$ and $\beta\neq 0$, by using a similar algorithm, we conclude that $$ u(x,y) = \iint\limits_\Omega {{K_{0\beta }}(x,y,{x_1},{y_1})f({x_1},{y_1})}d{x_1}d{y_1}, $$ where \begin{eqnarray*} {K_{0\beta }}(x,y,{x_1},{y_1}) &=& \theta (y)\left\{ {\theta ({y_1})\theta (x - {x_1})\left[ {G(x - {x_1},y,{y_1}) + {G_4}(x - {x_1},{x_1},y,{y_1})} \right]} \right. \hfill \\ && \left. { + \theta ( - {y_1})\theta (x - {\eta _1}){G_5}(x - {\eta _1},y,{\eta _1})} \right\} + \theta ( - y)\left\{ {\frac{{}}{{}}\theta ({y_1})\theta (\xi - {x_1}){G_3}(\xi ,{x_1},{y_1})} \right. \hfill \\ && \left. { + \theta ( - {y_1})\left[ {\theta (\xi - {\eta _1}){Q_1}(\xi ,\,{\eta _1}) + \frac{1}{2}\theta (\eta - {\eta _1})\theta ({\eta _1} - \xi )\theta (\xi - {\xi _1})} \right]} \right\}, \hfill \end{eqnarray*} with \begin{eqnarray*} {G_3}(x,{x_1},{y_1}) &=& \frac{1}{{\sqrt \pi }}\int\limits_0^{x - {x_1}} {\left[ {\frac{{{G_y}(z,{y_1},0)}}{{\sqrt z }} + \int\limits_0^z {\left\{ {\frac{{{G_{ys}}(s,{y_1},0)}}{{\sqrt {z - s} }} + R(z + {x_1},s + {x_1})\left[ {\frac{{{G_y}(s,{y_1},0)}}{{\sqrt s }}} \right.} \right.} } \right.} \hfill \\ && \left. {\left. {\left. { + \int\limits_0^s {\frac{{{G_{yt}}(t,{y_1},0)}}{{\sqrt {s - t} }}dt} } \right]} \right\}ds} \right]dz, \hfill \end{eqnarray*} \begin{eqnarray*} {Q_1}(x,{\eta _1}) &=& \frac{{2\beta }}{{\sqrt \pi }}\int\limits_0^{x - {\eta _1}} {\left[ {\frac{{Q({\eta _1},{\eta _1})}}{{\sqrt z }} + \int\limits_0^z {\left\{ {\frac{{{Q_s}(s,{\eta _1})}}{{\sqrt {z - s} }} + R(z + {\eta _1},s + {\eta _1})\left[ {\frac{{Q({\eta _1},{\eta _1})}}{{\sqrt s }}} \right.} \right.} } \right.} \hfill \\ && \left. {\left. {\left. { + \int\limits_0^s {\frac{{{Q_t}(t,{\eta _1})}}{{\sqrt {s - t} }}dt} } \right]} \right\}ds} \right]dz, \hfill \\ \end{eqnarray*} $$ {G_4}(t,{x_1},{y_1},y) = \int\limits_0^t {{G_3}(s,{x_1},{y_1}){G_y}(t - s,y,0)ds} $$ and $$ {G_5}(x,y,{\eta _1}) = \int\limits_0^x {{G_y}(x - {x_1},y,0){Q_1}({x_1},{\eta _1})d{x_1}} . $$ Thus we have partially proved the following lemma. \textbf{Lemma 1.} The unique regular solution of Problem B can be represented as follows $$ u\left( {x,y} \right) = \iint\limits_\Omega {K\left( {x,y,{x_1},{y_1}} \right)f\left( {{x_1},{y_1}} \right)d{x_1}d{y_1}},\,\,(x,y)\in \Omega,\eqno (25) $$ where $K\left( {x,y,{x_1},{y_1}} \right) \in {L_2}\left( {\Omega \times \Omega } \right)$ and $$K\left( {x,y,{x_1},{y_1}} \right) = {K_{\alpha \beta }}\left( {x,y,{x_1},{y_1}} \right), \quad \mbox{ if $\alpha \ne 0,$}$$ $$K\left( {x,y,{x_1},{y_1}} \right) = {K_{0\beta }}\left( {x,y,{x_1},{y_1}} \right), \quad \mbox{ if $\alpha = 0.$}$$ \verb"Proof": Expression (25) has been proved before. Let's see that $K\left( {x,y;{x_1},{y_1}} \right) \in {L_2}\left( {\Omega \times \Omega } \right)$. Note that in the kernel defined in (25), all the items are bounded except the first one. So, we only need to prove that $$ \theta \left( y \right)\theta \left( {{y_1}} \right)\theta \left( {x - {x_1}} \right)G\left( {x - {x_1},y,{y_1}} \right) \in {L_2}\left( {\Omega \times \Omega } \right). $$ From the representation of the Green's function $G\left( {x - {x_1},y,{y_1}} \right)$ given in (10), it follows that, for the aforementioned aim, it is enough to prove that (for $n = 0$): $$ B\left( {x - {x_1},y,{y_1}} \right) = \theta \left( y \right)\theta \left( {{y_1}} \right)\theta \left( {x - {x_1}} \right)\frac{1}{{2\sqrt {\pi \left( {x - {x_1}} \right)} }} \, \left[ {\exp \left\{ { - \frac{{{{\left( {y - {y_1}} \right)}^2}}}{{4\left( {x - {x_1}} \right)}}} \right\} - \exp \left\{ { - \frac{{{{\left( {y + {y_1}} \right)}^2}}}{{4\left( {x - {x_1}} \right)}}} \right\}} \right] $$ is bounded. First, note that $$ B\left( x - x_1,y,y_1 \right) \leq \frac{1}{2\sqrt {\pi \left( x - x_1 \right)} }e^{ - \frac{\left( y - y_1\right)^2}{4\left( x - x_1\right)}}. $$ Using this fact, we deduce that \begin{eqnarray*} \left\| B \right\|_{{L_2}\left( {\Omega \times \Omega } \right)}^2 &=& \int\limits_0^1 {dx\int\limits_0^1 {dy\int\limits_0^x {d{x_1}\int\limits_0^1 {{{\left| {B\left( {x - {x_1},y,{y_1}} \right)} \right|}^2}d{y_1} } } } }\\ & = & \int\limits_0^1 {dy} \int\limits_0^1 d{y_1}\int\limits_0^1 dx\int\limits_0^x {\left| B\left( x,y,{y_1} \right) \right|^2 d{x_1} } \\ &\leq & \int\limits_0^1 {dy} {\int\limits_0^1 {d{y_1}\int\limits_0^1 {\left| {B\left( {x,y,{y_1}} \right)} \right|^2} } }dx \leq \frac{1}{{4\pi }}\int\limits_0^1 {dy\int\limits_0^1 {d{y_1}} \int\limits_0^1 {\frac{1}{x}{e^{ - \frac{{{{\left( {y - {y_1}} \right)}^2}}}{{4x}}}}dx } } \\ &=& \frac{1}{4\pi }\int\limits_0^1{dy}\int\limits_0^1 {\frac{dx}{x}}\int\limits_0^1 {e^{ - \frac{\left(y -y_1\right)^2}{4x}}dy_1.} \end{eqnarray*} By means of the change of variables $\frac{y - y_1}{2\sqrt x } = y_2$, we get that this last expression is less than or equals to the following one $$ \frac{1}{{4\pi }}\int\limits_0^1 {dy} \int\limits_0^1 {\frac{{dx}}{x}} \int\limits_{\frac{{y - 1}}{{2\sqrt x }}}^{\frac{y}{{2\sqrt x }}} {{e^{ - y_2^2}}2\sqrt x d{y_2}} \leq \frac{1}{2\pi}\int\limits_0^1 {dy}\int\limits_0^1 {\frac{dx}{\sqrt {x} }}\int\limits_{-\infty}^{+\infty} {e^{-y_2}dy_2} = \frac{1}{\sqrt \pi}. $$ As consequence, $K\left( {x,y;{x_1},{y_1}} \right) \in {L_2}\left( {\Omega \times \Omega } \right)$ and Lemma 1 is completely proved. \hbox to 0pt{}\hfill$\rlap{$\sqcap$}\sqcup$ Define now $$ F_{\alpha\beta}(x)=\left\{ \begin{array}{l} F_1(x),\,\,\alpha\neq 0\\ F_2(x),\,\,\alpha=0.\\ \end{array} \right. $$ We have the following regularity result for this function: \textbf{Lemma 2.} If $f \in {C^1}(\overline{ \Omega} ),\,\,f(0,0) = 0$ and $Q \in {C^1}\left( {\left[ {0,1} \right] \times \left[ {0,1} \right]} \right)$, then $F_{\alpha \beta } \in {C^1}[0,1]$ and $F_{\alpha \beta }\left( 0 \right) = 0$. \verb"Proof:" Using the explicit form of the Green's function given in (10), it is not complicated to prove that function ${F_{\alpha \beta }}$, defined by formulas (16) and (19), belongs to the class of functions ${C^1}[0,1]$ and $F_{\alpha \beta }\left( 0 \right) = 0$. \\ Lemma 2 is proved. \hbox to 0pt{}\hfill$\rlap{$\sqcap$}\sqcup$ \textbf{Lemma 3.} Suppose that $Q \in {C^1}\left( {\left[ {0,1} \right] \times \left[ {0,1} \right]} \right)$ and $f \in {L_2}\left( \Omega \right)$, then ${F_{\alpha \beta }} \in {L_2}(\Omega )$ and $$ {\left\| {{F_{\alpha \beta }}} \right\|_{{L_2}\left( {0,1} \right)}} \leq C{\left\| f \right\|_{L_2(\Omega)}}.\eqno (26) $$ \verb"Proof:" Consider the following problem in ${\Omega _0}$: $$ {\omega _x} - {\omega _{yy}} = f(x,y),\,\,\,\,\,\,\,\,\,\,\,\,\,\,{\left. \omega \right|_{A{A_0} \cup {A_0}{B_0} \cup AB}} = 0.\eqno (27) $$ Obviously, we have that ${F_0}(x) = \mathop {\lim }\limits_{y \to 0} {\omega _y}(x,y)$. First, note that it is known \cite{mix} that problem (27) has a unique solution $\omega\in W_2^{1,2}({\Omega _0})$, and it satisfies the following inequality $$ \left\| \omega \right\|_{{L_2}({\Omega _0})}^2 + \left\| {{\omega _x}} \right\|_{{L_2}({\Omega _0})}^2 + \left\| {{\omega _y}} \right\|_{{L_2}({\Omega _0})}^2 + \left\| {{\omega _{yy}}} \right\|_{{L_2}({\Omega _0})}^2 \leq C\left\| f \right\|_{{L_2}({\Omega _0})}^2. \eqno (28) $$ Using now the obvious equality $$ {\omega _y}(x,0) = {\omega _y}(x,y) - \int\limits_0^y {{\omega _{yy}}(x,t)dt}, $$ we have that $$ \left\| {{\omega _y}(\cdot,0)} \right\|_{{L_2}(0,1)}^2 = {\int\limits_0^1 {\left| {{\omega _y}(x,0)} \right|^2} }dx = \int\limits_0^1 {dy{{\int\limits_0^1 {\left| {{\omega _y}(x,0)} \right|^2} }}dx} \leq C\left[ {\left\| {{\omega _y}} \right\|_{{L_2}({\Omega _0})}^2 + \left\| {{\omega _{yy}}} \right\|_{{L_2}({\Omega _0})}^2} \right].\eqno (29) $$ From (28) and (29) we obtain $$ {\left\| {{F_0}} \right\|_{{L_2}(0,1)}} = {\left\| {{\omega _y}\left( {\cdot,0} \right)} \right\|_{{L_2}(0,1)}} \leq C{\left\| f \right\|_{{L_2}({\Omega _0})}}.\eqno (30) $$ Now, by virtue of the conditions of Lemma 3 and the representations (16) and (19), from expression (30) and the Cauchy-Bunjakovskii inequalities, we get the estimate (26) and conclude the proof. \hbox to 0pt{}\hfill$\rlap{$\sqcap$}\sqcup$ Denote now $\| \cdot \|_l$ as the norm of the Sobolev space $H^l(\Omega)\equiv W_2^l(\Omega)$ with $W_2^0(\Omega)\equiv L_2(\Omega)$. \textbf{Lemma 4.} Let $u$ be the unique regular solution of Problem B. Then the following estimate holds: $$ {\left\| u \right\|_1} \leq c{\left\| f \right\|_0}.\eqno (31) $$ Here $c$ is a positive constant that does not depend on $u$. \verb"Proof:" By virtue of Lemma 3, and from (20) and (21), we deduce that $$ {\left\| {\tau '} \right\|_{{L_2}\left( {0,1} \right)}} \leq C{\left\| {{F_{\alpha \beta }}} \right\|_{{L_2}\left( {0,1} \right)}} \leq C{\left\| f \right\|_0}. $$ The result follows from expression (22). \hbox to 0pt{}\hfill$\rlap{$\sqcap$}\sqcup$ \textbf{Definition 1}. We define the set $W$ as the set of all the regular solutions of Problem B. A function $u \in {L_2}\left( \Omega \right)$ is said to be a strong solution of Problem B, if there exists a functional sequence $\left\{ {{u_n}} \right\} \subset W$, such that ${u_n}$ and $L{u_n}$ converge in ${L_2}\left( \Omega \right)$ to $u$ and $f$ respectively. Define $\mathbb{L}$ as the closure of the differential operator $\mathbb{L}:\,W\rightarrow {L_2}\left( \Omega \right)$, given by expression (2). Note that, according to the definition of the strong solution, the function $u$ will be a strong solution of Problem B if and only if $u \in D\left( \mathbb{L} \right)$. Now we are in a position to prove the following uniqueness result for strong solutions. \textbf{Theorem 2.} For any function $Q \in {C^1}\left( {\left[ {0,1} \right] \times \left[ {0,1} \right]} \right)$ and $f \in {L_2}\left( \Omega \right)$, there exists a unique strong solution $u$ of Problem B. Moreover $u\in W_2^1\left( \Omega \right) \cap W_{x,y}^{1,2}\left( {{\Omega _1}} \right) \cap C\left( {\overline \Omega } \right)$, satisfies inequality (31) and it is given by the expression (25). \verb"Proof:" Let $C_{0}^{1}\left( {\overline{\Omega }} \right)$ be the set of the $C^1\left(\overline{\Omega}\right)$ functions that vanish in a neighborhood of $\partial \Omega$ ($\partial \Omega$ is a boundary of the domain $\Omega$). Since $C_{0}^{1}\left( {\overline{\Omega }} \right)$ is dense in ${L_2}\left( \Omega \right)$, we have that for any function $f \in {L_2}\left( \Omega \right)$, there exist a functional sequence ${{f}_{n}}\in C_{0}^{1}\left( {\overline{\Omega }} \right)$, such that $\left\| {{f_n} - f} \right\| \to 0$, as $n \to \infty$. It is not difficult to verify that if $f_n\in C_0^1\left( {\overline{\Omega }} \right)$ then $F_{\alpha \beta n}\in C^1(\left[ 0,1 \right])$ (with obvious notation). Therefore equations (14) and (18) can be considered as a second kind Volterra integral equations in the space $C^1(\left[ 0,1 \right])$. Consequently, we have that ${\tau '_n}\left( x \right) = {u_{nx}}\left( {x,0} \right) \in {C^1}\left[ {0;1} \right]$. Due to the properties of the solutions of the boundary value problem for the heat equation in $\Omega_0$ and the Darboux problem, by using the representations (6) and (9), we conclude that ${u_n} \in W$ for all $f_n\in C_{0}^{1}\left( {\overline{\Omega }} \right)$. By virtue of the inequality (31) we get $$ {\left\| {{u_n} - u} \right\|_1} \leq c{\left\| {{f_n} - f} \right\|_0} \to 0. $$ Consequently, $\left\{ {{u_n}} \right\}$ is a sequence of strong solutions, hence, Problem B is strongly solvable for all right hand $f\in L_2(\Omega)$, and the strong solution belongs to the space $W_2^1\left( \Omega \right) \cap W_{x,y}^{1,2}\left( {{\Omega _1}} \right) \cap C\left( {\overline \Omega } \right)$. Thus, Theorem 2 is proved. \hbox to 0pt{}\hfill$\rlap{$\sqcap$}\sqcup$ Consider now, for all $n=2,3, \ldots$, the sequence of kernels given by the recurrence formula $$ {K_n}(x,y;\,{x_1},{y_1}) = \iint\limits_\Omega {K(x,y;\,{x_2},{y_2}){K_{(n - 1)}}({x_2},{y_2},{x_1},{y_1})d{x_2}d{y_2}\,\,}, $$ with $$ {K_1}(x,y;\,{x_1},{y_1}) = K(x,y;\,{x_1},{y_1}), $$ and $K$ defined in Lemma 1. \textbf{Lemma 5. } For the iterated kernels ${K_n}(x,y;\,{x_1},{y_1})$ we have the following estimate: $$ \left| {{K_n}(x,y;\,{x_1},{y_1})} \right| \leq {(\sqrt {\pi \,} M)^n}{\left( {\frac{3}{2}} \right)^{n - 1}}\,\frac{{{{(x - {x_1})}^{\frac{n}{2} - 1}}}}{{\Gamma \left( {\frac{n}{2}} \right)}}\,\,,\,\,\,\,\,n = 1,2,3...,\eqno (32) $$ where $ M=\underset{\begin{smallmatrix} \left( x,y \right)\in \Omega \\ \left( {{x}_{1}},{{y}_{1}} \right)\in \Omega \end{smallmatrix}}{\mathop{\max }}\,\left| \sqrt{x-{{x}_{1}}}\,K\left( x,y;\,{{x}_{1}},{{y}_{1}} \right) \right| $ and $\Gamma$ is the Gamma-function of Euler. \verb"Proof:" The proof will be done by induction in $n$. Taking the representation of the Green's function given in (10) into account, and from the representation of the kernel $K(x,y;\,{x_1},{y_1})$ at $n = 1$, the inequality (32) $$ \left| {{K_1}(x,y;\,{x_1},{y_1})} \right| \leq M{(x - {x_1})^{ - \frac{1}{2}}} $$ is automatically deduced. Let (32) be valid for $n = k - 1$. We will prove the validity of this formula for $n = k$. To this end, by using inequality (32), at $n = 1$ and $n = k - 1$, we have that \begin{eqnarray*} \left| K_k(x,y;x_1,y_1) \right| &=& \left| \iint\limits_\Omega K(x,y;x_2,y_2)K_{(k - 1)}(x_2,y_2,x_1,y_1)dx_2dy_2 \right| \hfill \\ & \leq & \iint\limits_\Omega {\left| K(x,y;x_2,y_2) \right| \, \left| K_{(k - 1)}(x_2,y_2;x_1,y_1) \right|dx_2dy_2} \hfill \\ & \leq & \iint\limits_\Omega {\theta (x - x_2)M(x - x_2)^{ - \frac{1}{2}}} \, \theta (x_2 - x_1) \, (\sqrt \pi {M})^{k - 1}\left(\frac{3}{2}\right)^{k - 2} \frac{(x_2 - x_1)^{\frac{k}{2} - \frac{3}{2}}}{\Gamma (\frac{k - 1}{2})}dx_2dy_2 \hfill \\ & \leq & {M^k}{(\sqrt \pi )^{k - 1}}{\left(\frac{3}{2}\right)^{k - 2}} \, \frac{1}{{\Gamma (\frac{{k - 1}}{2})}}\int\limits_{{x_1}}^x {d{x_2}\int\limits_{ - \frac{1}{2}}^1 {{{(x - {x_2})}^{ - \frac{1}{2}}}{{({x_2} - {x_1})}^{\frac{k}{2} - \frac{3}{2}}}dy_2.} } \end{eqnarray*} Evaluating the previous integrals we have that \begin{eqnarray*} \left| {{K_k}(x,y;{x_1},{y_1})} \right| & \leq & {M^k}{(\sqrt \pi )^{k - 1}}{\left(\frac{3}{2}\right)^{k - 1}}\frac{{{{(x - {x_1})}^{\frac{k}{2} - 1}}}}{{\Gamma (\frac{{k - 1}}{2})}} \int\limits_0^1 {{\sigma ^{ - \frac{1}{2}}}{{(1 - \sigma )}^{\frac{k}{2} - \frac{3}{2}}}d\sigma} \\ & =&{ (\sqrt \pi } M{)^k}{\left(\frac{3}{2}\right)^{k - 1}}\frac{{{{(x - {x_1})}^{\frac{k}{2} - 1}}}}{{\Gamma (\frac{k}{2})}}, \hfill \\ \end{eqnarray*} which proves Lemma 5. \hbox to 0pt{}\hfill$\rlap{$\sqcap$}\sqcup$ Now we are in a position to prove the final result of this paper, which gives us the Volterra property for the inverse of operator $\mathbb{L}$. \textbf{Theorem 3.} The integral operator defined in the right hand of (25), i.e. $$ {\mathbb{L}^{ - 1}}f(x,y) = \iint\limits_\Omega {K(x,y;\,{x_1},{y_1})f({x_1},{y_1})d{x_1}d{y_1},}\eqno (33) $$ has the Volterra property (it is almost continuous and quasi-nilpotent) in ${L_2}$($\Omega $). \verb"Proof:" Since the continuity of this operator follows from the fact that $K \in {L_2}(\Omega \times \Omega )$. To prove this theorem, we only need to verify that operator ${\mathbb{L}^{ - 1}}$, defined by (33), is quasi-nilpotent, i.e. $$ \mathop {\ell im}\limits_{n \to \infty } \left\| {{\mathbb{L}^{ - n}}} \right\|_0^{\frac{1}{n}} = 0,\eqno (34) $$ where $$ {\mathbb{L}^{ - n}} = {\mathbb{L}^{ - 1}}\left[ {{\mathbb{L}^{ - (n - 1)}}} \right]\,\,\,\,\,\,,\,\,\,\,n = 1,2,3, \ldots $$ From (33), and by direct calculations, one can easily arrive at the following expression: $$ {\mathbb{L}^{ - n}}f(x,y) = \iint\limits_\Omega {{K_n}(x,y;\,{x_1},{y_1})f({x_1},{y_1})d{x_1}d{y_1}.}\eqno (35) $$ Consequently, using the inequality of Schwarz and expression (32), from the representation (35) we obtain that \begin{eqnarray*} \left\| {{\mathbb{L}^{ - n}}f} \right\|_0^2 &=& \iint\limits_\Omega {{{\left| {{\mathbb{L}^{ - n}}f} \right|}^2}dxdy} = \iint\limits_\Omega {{{\left[ {\iint\limits_\Omega {{K_n}(x,y;{x_1},{y_1})f({x_1},{y_1})d{x_1}d{y_1}}} \right]}^2}dxdy } \hfill \\ & \leq & \iint\limits_\Omega \left[ {\left( {\iint\limits_\Omega {{{\left| {f\left( {{x_1},{y_1}} \right)} \right|}^2}d{x_1}d{y_1}}} \right)\,\left( {\iint\limits_\Omega {{{\left| {{K_n}\left( {x,y;\,{x_1},{y_1}} \right)} \right|}^2}d{x_1}d{y_1}}} \right)} \right]dxdy \\ & \leq & \left( {\frac{3}{2}\sqrt \pi M} \right)^{2n}\frac{1}{{n\left( {n - 1} \right){\Gamma ^2}\left( {\frac{n}{2}} \right)}}\left\| f \right\|_0^2. \end{eqnarray*} From here we get $$ {\left\| {{\mathbb{L}^{ - n}}} \right\|_0} \leq \left( {\frac{3}{2}\sqrt \pi M} \right)^{n}\frac{1}{{\Gamma (1 + \frac{n}{2})}}. $$ From the last equality one can state the validity of the equality (34) and Theorem 3 is proved.\hbox to 0pt{}\hfill$\rlap{$\sqcap$}\sqcup$ \textbf{Consequence 1.} Problem B has the Volterra property. \textbf{Consequence 2.} For any complex number $\lambda$, the equation $$ \mathbb{L}u - \lambda u = f\eqno (36) $$ is uniquely solvable for all $f \in {L_2}(\Omega )$. Due to the invertibility of the operator $\mathbb{L}$, the unique solvability of the Eq.(36) is equivalent to the uniqueness of solution of the equation $$ u - \lambda {\mathbb{L}^{ - 1}}u = {\mathbb{L}^{ - 1}}f, $$ which is a second kind Volterra equation. This proves Consequence 2 of Theorem 3. \section*{Competing interests} The authors declare that they have no competing interests. \section*{Authors' contributions} The four authors have participated into the obtained results. The collaboration of each one cannot be separated in different parts of the paper. All of them have made substantial contributions to the theoretical results. The four authors have been involved in drafting the manuscript and revising it critically for important intellectual content. All authors have given final approval of the version to be published. \section*{Acknowledgement} This research was partially supported by Ministerio de Ciencia e Innovaci\'{o}n-SPAIN, and FEDER, project MTM2010-15314 and KazNPU Rector's grant for 2013 .
\section{Introduction} \label{sec:1 Wireless networks, due to their broadcast nature, are vulnerable to being overheard, and hence security is a primary concern. The standard method of providing security against eavesdroppers is to encrypt the information so that it is beyond the eavesdropper's computational capabilities to decrypt the message \cite{stinson2006cryptography}; however, the vulnerability shown by many implemented cryptographic schemes, the lack of a fundamental proof establishing the difficulty of the problem presented to the adversary, and the potential for transformative changes in computing motivate forms of security that are provably everlasting. In particular, when a cryptographic scheme is employed, the adversary can record the clean cypher and recover it later when the cryptographic algorithm is broken \cite{bensonverona}, which is not acceptable in sensitive applications requiring everlasting secrecy. The desire for such everlasting security motivates considering emerging information-theoretic approaches, where the eavesdropper is unable to extract from the received signal any information about the secret message. In 1949, Shannon introduced information-theoretic, or perfect, secrecy \cite{shannon1949communication}. If the uncertainty of the message after seeing the cypher is equal to the uncertainty of the message before seeing the cypher, we have perfect secrecy without any condition on the eavesdropper's capabilities. Wyner later showed that if the eavesdropper's channel is degraded with respect to the main channel, adding some randomness to the codebook allows the achievement of a positive secrecy rate \cite{wyner1975wire}. Csisz{\'a}r and K{\"o}rner extended the idea to more general cases, where the eavesdropper's channel is not necessarily degraded with respect to the main channel, but it must be ``more noisy'' or ``less capable'' than the main channel \cite{csiszar1978broadcast}. When such an advantage does not exist, one can turn to approaches based on ``public discussion'' \cite{maurer1993secret,ahlswede1993common}, but these approaches, while they could be used to generate an information-theoretically secure one-time pad, are generally envisioned for secret key agreement to support a cryptographic approach \cite[Chapter 7.4]{bloch2011physical} rather than one-way secret communication. We will show later the relation between our proposed scheme and public discussion, noting, in particular, that the proposed scheme can be used in conjunction with public discussion when appropriate. Consequently, the desirable situation for achieving information-theoretic secrecy is to have a better channel from the transmitter to the intended receiver than that from the transmitter to the eavesdropper. However, this is not always guaranteed, particularly in wireless systems where the eavesdropper can have a large advantage over the intended receiver. In the case of a passive adversary, the eavesdropper can be very close to the transmitter or it can use a directional antenna to improve its received signal, while there is often no way for the legitimate nodes to know the eavesdropper's location or its channel state information. Recent authors have considered approaches that relax the need for assumptions on Eve's location or channel in one-way systems. For cases when the eavesdropper location is unknown (which means the case of a ``near Eve'' must be considered), approaches largely based on the cooperative jamming approach of \cite{negi2005secret} and \cite{goel2005secret} have been considered \cite{he2008two,lai2007cooperative}. However, all of these approaches require either multiple antennas, helper nodes, and/or fading (for example, \cite{gopala2008secrecy,Dennis2011JSAC,sheikholeslami2012physical}), and many are susceptible to attacks such as pointing directive antennas at one or both communicating parties. For a one-way scenario with a single antenna where Bob's channel is worse than Eve's, Cachin and Maurer \cite{cachin1997unconditional} exploited the realizability of hardware to consider the case of everlasting security, as is our interest. In particular, they introduced the ``bounded storage model'' in which the receiver cannot store the information it would need to eventually break the cypher. This novel approach suffers from two shortcomings: (1) by Moore's Law (see NAND scaling plot at \cite{kuchibhatla2010imft}), the density of memories increases at an exponential rate; (2) memories can be stacked arbitrarily subject only to (very) large space limitations. Hence, although the bounded storage model is a viable approach to everlasting security, it is difficult to pick a memory size beyond which it will be effective, making its employment for secret wireless communication difficult. Rather than attacking the memory in the receiver back-end, our contention is that one should instead consider attacking the receiver front-end and analog-to-digital (A/D) conversion process, where technology progresses slowly and there exist well-known techniques for severely handicapping the component. And, unlike memory, A/D's cannot be stacked arbitrarily, as clock jitter prevents the timing required for bit detection; in fact, high-quality A/D's already employ parallelization to the limit of the jitter. And, importantly from a long-term perspective, there is a fundamental bound on the ability to perform A/D conversion \cite{krone2009fundamental,krone2010fundamental}. \begin{figure} \begin{center} \includegraphics[width=.5\textwidth]{wiretap_channel2.eps} \end{center} \caption{The message $X$ is observed at Bob and Eve through the transmitter, the AWGN channels with different noise variances, and their respective receivers with (possibly nonlinear) functions $g(.)$, $f(.)$, and $f_E(.)$. The sequence $\underline{k}$ is a cryptographic key shared by Alice and Bob, which is assumed to be obtained by Eve immediately \textit{after} she has recorded $Z$.} \label{fig:p} \end{figure} Consider the channel model shown in Figure \ref{fig:p}, which reflects the understanding that in an adversarial game in modern communication systems, it is the interference effects on wideband receiver front-ends rather than the baseband processing that is the significant detriment \cite{Hashemi2011RFIC}. In particular, the signal is subject to a variety of distortions due to the RF front-end of the receiver and the analog-to-digital conversion. A large interferer, even if it is orthogonal to the signal of interest and thus (supposedly) easily rejected by baseband processing, can saturate the receiver front-end, leading to nonlinearities, and, of particular interest here, reducing the receiver's dynamic range (i.e. resolution) significantly. The primary focus of this paper is to exploit the receiver processing effects for security. In particular, based on a pre-shared key between Alice and Bob that only needs to be kept secret for the duration of the wireless transmission (i.e. it can be given to Eve immediately afterward), we consider how inserting intentional (but known to Bob) distortion on the transmitted signal can provide information-theoretic security. In particular, since Bob knows the distortion, he can undo its effect before his A/D, whereas Eve must store the signal and try to compensate for the distortion after her A/D. Since the A/D is necessarily non-linear, the operations are not necessarily commutative and there is the potential for information-theoretic security. This paper introduces this idea and initiates its investigation. As a first example, we perform a rapid power modulation between two vastly different power levels at the transmitter and put the reciprocal of that power gain before Bob's A/D. In particular, cellular (and other) networks usually have significantly more power available for users at many locations than their lowest data rate requires for successful transmission. For example, users near a base station in a cellular system have the capability to transmit significantly more power than the minimum required to convey a high-quality voice signal. Hence, a secure communication system to cover a restricted area (e.g. a company's building) built on analogous link budgets to cellular technology would have the capability to transmit excess power to enable secure communication, as follows. Suppose Alice employs an ephemeral cryptographic key known only to her and Bob to rapidly modulate her transmit power between the minimum required for successful transmission and the maximum available from her radio. This power modulation can be done quite rapidly, as modern power amplifiers can easily have their power switched at high bandwidths \cite{kahn1952single}\cite[Chapter~7]{kenington2000high}. Bob, since he knows the key, places a gain before his A/D that changes rapidly in concert with the transmitted power to ensure that the received signal is matched to the range of the A/D. Since the power can be changed every symbol, Eve cannot use any type of automatic gain control (AGC) loop and is left trying to select a gain that trades off resolution and the probability of overflow of her A/D. By exploiting the resulting distortion, information-theoretic secrecy can be obtained, even if Eve is given the key immediately after message transmission. The rest of paper is as follows. Section \ref{sec:2} describes the system model, metrics, and the proposed idea in detail. In Section \ref{sec:3}, the proposed method is applied to settings with noisy channels and noiseless channels, respectively, to find achievable secrecy rates in each case, and an asymptotic analysis of the proposed method is provided. In Section \ref{sec:4}, the results of numerical examples for various realizations of the system are presented. Conclusions and ideas for future work are discussed in Section \ref{sec:7}. \section{System Model and Approach}\label{sec:2 \subsection{System Model and Metric} We consider a simple wiretap channel, which consists of a transmitter, Alice, a receiver, Bob, and an eavesdropper, Eve. Eve is a passive eavesdropper, i.e. she just tries to obtain as much information as possible to recover the message that Alice sends and she does not attempt to actively thwart (i.e. via jamming, signal insertion) the legitimate nodes. Therefore, the location and channel state information of Eve can be difficult to obtain and thus is assumed unknown to the legitimate nodes. We assume that Alice and Bob either pre-share a (very) short initial key or that they employ a standard key agreement scheme (e.g. Diffie-Hellman \cite{diffie1976new}, which is very efficient in passive environments) to generate a shared key. This initial key will be used to generate a very long key-sequence by using a standard cryptographic method such as AES in counter mode (CTR). Considering the fact that for each $2^{38}$ bits of the key-sequence, a 96-bit new initial vector (IV) or a 128-bit new initial key must be sent from Alice to Bob \cite[Chapter 5]{paar2010understanding}, the secrecy rate overhead that this key (or IV) exchange imposes is at most $128/2^{38}=2^{-29}$, which is negligible. Another method is to use standard methods that are specifically designed for generating stream-ciphers, such as Trivium (more methods can be found in \cite{robshaw2008new}), which can generate $2^{64}$ bits of key-sequence for a 80-bit key and a 80-bit IV. Thus, the rate overhead that Trivium places on our scheme will be $80/2^{64}<2^{-55}$, which is negligible. By using these cryptographic algorithms to perform key-expansion, we assume that Eve cannot recover the initial key before the key renewal and during the transmission period, i.e. we assume that the computational power of Eve during the time of transmission is not unlimited. However, our system design only employs the key ephemerally. In fact, we assume (pessimistically) that Eve is handed the full key (and not just the initial key) as soon as transmission is complete. Hence, unlike cryptography, even if the encryption system is broken later, the eavesdropper obtains access to an unlimited computational power, or other forms of computation such as quantum computers are implemented, Eve will not have enough information to recover the secret message. We consider a memoryless one-way communication system, and assume that both Bob and Eve are at a unit distance from the transmitter by including variations in the path-loss in the noise variance. Thus, the channel gain of both channels is unity and both channels experience additive white Gaussian noise (AWGN). Let $n_B$ and $n_E$ denote the zero-mean noise processes at Bob's and Eve's receivers with variances $\sigma_B^2$ and ${\sigma_E}^2$, respectively. Let $\hat{X}$ denote the input of both channels, $\hat{Y}$ denote the received signal at Bob's receiver, and $\hat{Z}$ denote the received signal at Eve's receiver. The signal at Bob's receiver is: \[\hat{Y}=\hat{X}+n_B,\] and the signal at Eve's receiver is: \[\hat{Z}=\hat{X}+n_E.\] We assume that location of Alice is known to Eve. Also, Alice knows either Eve's location, or in the case that she does not know Eve's location, she sets a value that works over a set of locations (for example, the minimum possible distance between Alice and Eve). If the location of Eve is completely unknown, Eve's distance can assumed to be zero and, as will be shown in Section \ref{sec:4}, the legitimate nodes will still be able to obtain a positive secrecy rate by using the proposed scheme. Both Bob and Eve employ high precision uniform analog-to-digital converters. The effect of the A/D on the received signal (quantization error) is modeled by a quantization noise due to the limitation in the size of each quantization level, and a clipping function due to the quantizer's overflow. The quantization noise in this case is (approximately) uniformly distributed \cite{widrow2008quantization}, so we will assume it is uniformly distributed throughout the paper. For an \textit{m}-bit quantizer ($b=2^m$ gray levels) over the full dynamic range $[-l,l]$, two adjacent quantization levels are spaced by $\delta={2l}/{b}$, and thus the quantization noise is uniformly distributed over an interval of length $\delta$. Quantizer overflow happens when the amplitude of the received signal is greater than the quantizer's dynamic range, which can be modeled by a clipping function. We assume that Alice knows an upper bound on Eve's current A/D conversion ability (without any assumption on Eve's future A/D conversion capabilities). Let $X$ denote the current code symbol, which we assume is taken from a standard Gaussian codebook where each entry has variance $P$, i.e. $X\sim \mathcal{N}(0,P)$. Note that although the Gaussian codebook is optimal to achieve the secrecy capacity in the case of AWGN wiretap channels, because we consider quantization errors in our model, the Gaussian codebook is no longer optimum, implying that our results represent achievable rates but not upper bounds. From \cite{bloch2008secrecy}, for an arbitrary stationary memoryless wiretap channel with arbitrary input and output alphabets, any secrecy rate \[ \hat{R}_s< \max_{X\rightarrow YZ} [I(X;Y)-I(X;Z)] \] is achievable. Now, we define the following max-min criteria: \begin{equation} R_s=\max_{s\in\mathcal{S}}\min_{s'\in \mathcal{S}'} \hat{R}_s(s,s') \label{eq:101} \end{equation} where $\mathcal{S}'$ is the set of strategies that Eve can take during transmission, and $\mathcal{S}$ is the set of strategies that Alice can take. Eve's problem is to find a strategy, $s'\in\mathcal{S}'$, to modify her channel to minimize the secrecy rate. On the other hand, Alice's problem is to find a strategy, $s\in\mathcal{S}$, to modify the transmit signal to maximize this worst-case secrecy rate. When cryptographic key expansion schemes are employed, the key-sequence is not quite memoryless. But, based on the assumption that Eve cannot restrict the rest of the key sequence based on the observed symbols, we assume independence. Hence, although in general the strategy taken by Eve is not memoryless, here considering strategies with memory does not help her to increase the information-leakage; thus, we restrict $\mathcal{S}'$ to memoryless strategies. Further, we give the key to Eve after completion of the transmission and show she cannot recover the lost information she would need to obtain the secret message from the recorded symbols. \subsection{General Nonlinearity: Rough Analysis Our goal is to consider how Alice and Bob can employ bits of the shared key to modify their radios as shown in Figure \ref{fig:p} to gain (or maximize) an information-theoretic advantage. For now, assume that they insert general memoryless nonlinearities $g(.)$ at the transmitter and $f(.)=g^{-1}(.)$ at the receiver based on the key. Suppose that Eve is able to obtain the key just after the transmission is finished; considering for the moment that she applies $g^{-1}(.)$ to $Z$, one sees how the security is (potentially) obtained: Bob sees $g(X)$ through $g^{-1}(.)$ and the A/D, whereas Eve sees those operations in reverse. Since nonlinear operations are not (necessarily) commutative, the signals are not the same and there is the potential for some form of information-theoretic security. Now, stepping back to allow Eve to use the long key sequence, $\underline{k}$, in whatever manner she wants after she has recorded the transmission yields an illustrative information-theoretic model. In particular, using the same random coding arguments as for fading channels, consider a collection of functions $\mathcal{G}$, from which $\underline{k}$ selects a function $g(.)$ for each transmitted symbol; then, the secrecy rate is: \[ R_s=E_{g(.)}[I(X;Y|g(.))-I(X;Z|g(.))] \] Let us be pessimistic and assume $\sigma_E^2=0$. Furthermore, to get some insight, assume temporarily that $\sigma_B^2=0$, corresponding to a short-range situation which is not power-limited. For $\sigma_B^2=0$, $Y$ does not depend on $\underline{k}$ and thus using the approach for analyzing quantizers of \cite[pg. 251]{cover2006elements}, which is accurate at high resolution: \begin{align*} &R_s=E_{g(.)}[I(X;Y)-I(X;Z|g(.))]\\ &=E_{g(.)}[H(Y)-H(Y|X)-(H(Z|g(.))-H(Z|X,g(.)))]\\ &\approx E_{g(.)}[h(\tilde{Y})-\log(\delta)-(h(\tilde{Z}|g(.))-\log(\delta))]\\ &= E_{g(.)}[h(\tilde{Y})-h(\tilde{Z}|g(.))]\\ &=E_{g(.)}[h(X)-h({g(X)})] \end{align*} where $\tilde{Y}$ and $\tilde{Z}$ are the inputs to Bob and Eve's A/D converters, respectively. It then becomes apparent that the gain observed here for high-resolution A/D's at both Bob and Eve is a shaping gain between $X$ and $g(X)$. Whereas we think of shaping gains as tending to be relatively small (1.53 dB on the Gaussian channel \cite{cald1990}), that is because the generally considered gains are between the optimal (Gaussian) shaping and a standard but reasonable (uniform) shaping. In our design scenario, if we are able to severely distort the signal, the gains can become enormous. We quickly caveat this conclusion by noting that the assumption $\sigma_B^2=0$ is critical, since those $g(.)$ which are most distorting can also cause significant ``noise enhancement'' on the channel from Alice to Bob. Hence, unless the noise is truly negligible (i.e. very short range communication), judgment should be reserved on the applicability of the technique until $\sigma_B^2\neq 0$ is considered in Section \ref{sec:3}. \subsection{Rapid power modulation for secrecy} \begin{figure} \begin{center} \includegraphics[width=.5\textwidth]{wiretap_channel.eps} \end{center} \caption{Alice and Bob share a cryptographic key that determines the value of $A$ at each time instance. Eve puts a (possibly variable) gain before her A/D to decrease the A/D erasures and/or overflows and hence increase the information leakage.} \label{fig:p10} \end{figure} For the rest of the paper, we simplify the operator $g(.)$ to a random gain to consider a practical architecture easily implemented and discuss specific operating scenarios. Our goal is to achieve a positive secrecy rate by confusing Eve's A/D. Throughout this paper we assume that Eve is able to employ just one A/D, and Eve with multiple A/D's is briefly discussed in Section \ref{sec:7}. The random gain is from a fixed probability distribution and is multiplied to the signal amplitude of each symbol that Alice transmits. Suppose that $A$ denotes the random variable associated with this random gain, and the probability density function (pdf) of this gain is $p_A(a)$ where $a \in \mathcal{A}$ (see Figure \ref{fig:p10}). The pdf of $A$ is known to all nodes, but only legitimate nodes know the exact sequence of values of $A$ (i.e. $ a_1,a_2,a_3,\cdots$) that is applied to the symbol sequence. We want to find a probability distribution for $A$ that maximizes this secrecy rate such that it does not change the average power of the transmitted signal, i.e. $\text{E}[|A|^2]=1$. To control the number of key bits required, we consider that $|A|$ is drawn from one of two levels $A_1$ and $A_2$ with random polarity (i.e. $\mathcal{A}=\{A_1,-A_1,A_2,-A_2\}$): \[ Pr(A=a)=\left\{ \begin{array}{l l} p, & \quad a=A_{1}\\ 1-p,& \quad a=A_{2}\\ \end{array} \right. \] and $Pr\{A>0\}=Pr\{A<0\}=\nicefrac{1}{2}$. Suppose that $A_1$ is the large gain and $A_2$ is the small gain that the transmitter applies and denote the ratio between them $r=\frac{A_1}{A_2}$. Since Bob shares the (long) key with Alice, he easily ``inverts'' the gain $A$ to operate his A/D properly, whereas Eve will struggle with such. In essence, we are inducing a fading channel at Bob that he is able to equalize \textit{before} his A/D, whereas Eve cannot. Bob applies the reciprocal of $A$ before his A/D and thus given $A$, the signal that Bob's A/D sees is: \begin{equation} \tilde {Y}=X+\frac{n_B}{A} \label{eq:1a} \end{equation} To cancel the effect of this gain, Eve also applies an arbitrary (possibly random) gain, $1/G$. So, the signal at Eve's A/D given $A$ and $G$ is: \begin{equation} \tilde {Z}=\frac{A}{G}X+\frac{n_E}{G} \label{eq:1b} \end{equation} Suppose that Eve knows the pdf of $A$; hence, she tries to find a probability density function $p_G(g)$ for $G$ such that it minimizes the secrecy rate $R_s$. On the other hand, Alice sets the pdf parameters such that no matter what $p_G(g)$ Eve chooses, some secrecy rate $R_s$ is always guaranteed. Hence, the maxi-min criteria in (\ref{eq:101}) turns into: \begin{equation} R_s=\max_{p, A_1, A_2}\min_{p_G(.)} \hat{R}_s(p_G(.),A_1,A_2,p)\label{eq:10} \end{equation} Obviously, larger $r=\frac{A_1}{A_2}$ leads to more eavesdropper confusion. However, because $E[|A|^2]=1$, $r \gg 1$ leads to a small $A_2$, and Bob then suffers noise enhancement. We talk about the choice of $r$ in the next paragraph. Recall the potential operating scenario from Section \ref{sec:1}, and assume that system radios are operating in a scenario where they have adequate power amplifier headroom, as in the ``near'' situation in cellular systems \cite{kohno1995spread}, and the user's noise is relatively negligible. However, an Eve at the same range can also intercept the signal. By changing the power of the transmitters between the power-controlled level (e.g. $A_2$), where it meets the receiver requirements and its maximum power (e.g. $A_1$), Bob, knowing the sequence, obtains a signal that is at least equivalent to operating at its power controlled level and thus sees little degradation in information transmission. The ratio between the large gain and the small gain, $r$, can be chosen such that in the case of $A=A_2$ (small gain), the minimum acceptable signal level at Bob's receiver is satisfied. On the other hand, Eve's A/D struggles even to record a reasonable form of the signal; hence, she sees significant degradation, and information-theoretic security is obtained. Also, because the power level is changed very fast (at every symbol), the automatic gain control (AGC) at the eavesdropper's receiver cannot follow the deep fades that cause erasures and/or strong signals that cause A/D saturation. To choose optimum values for $A_1$, $A_2$, and $p$, note that the following constraints must be met: \begin{equation} \frac{A_1}{A_2}=r\quad \text{and}\quad p A_1^2+(1-p) A_2^2=1 \label{eq:cons} \end{equation} Hence, two of these values are constrained by the system parameter $r$ and conservation of transmission power, and the transmitter is free to choose only one (e.g. $p$). Thus, equation (\ref{eq:10}) reduces to: \begin{equation} R_s=\max_{p}\min_{p_G(.)} \hat{R}_s(p_G(.),p)\label{eq:11} \end{equation} Eve can employ a number of countermeasures to decrease $R_s$. She can find an optimum probability density function that minimizes $R_s$, or she can employ a better A/D to decrease erasures and/or overflows of her A/D. In the sequel, we will consider these scenarios and examine the secrecy rate $R_s$ that can be achieved by the proposed method in each case. \section{Achievable Secrecy Rates}\label{sec:3} In this section the secrecy rates that can be achieved considering the non-idealities of the A/D's at the front-ends of Bob and Eve's receivers are studied. In the first part, the channel between Alice and Bob and the channel between Alice and Eve are considered to be AWGN channels. In the second part, to get more insight into the problem, the noise is removed from the channels and only the effect of A/D's on the signals will be considered. \subsection{Noisy channels} Consider the derivation of $I(X;Y|A=a)-I(X;Z|A=a,G=g)$. Clearly, each of $h(Y|A=a)$, $h(Y|X,A=a)$, $h(Z|A=a,G=g)$, and $h(Z|X,A=a,G=g)$ are required. Since for given gains at Alice and Eve, i.e. $A=a$ and $G=g$, by substituting $Z$ with $Y$ and $g$ with $a$ (Figure \ref{fig:p10}), the equations for $h(Y|A=a)$, $h(Y|X,A=a)$ can be derived from the equations for $h(Z|A=a,G=g)$ and $h(Z|X,A=a,G=g)$, we just show the calculations for the latter here. In this section all the mutual information, entropy, and probability density functions are calculated given that $A=a$ and $G=g$. Recall that throughout this paper the non-idealities of the A/D's are modeled by an additive uniformly distributed quantization noise and a clipping function; hence,the signal at the output of Eve's A/D is: \[ Z=\left\{ \begin{array}{l l} \tilde{Z}+n_q, & \quad |\tilde{Z}|<l\\ +l, & \quad \tilde{Z}>l\\ -l, & \quad \tilde{Z}<-l\\ \end{array} \right. \] where $\tilde{Z}=\frac{aX}{g}+\frac{n_E}{g}$ and $l$ is determined by the range $[-l,l]$ of the A/D. Thus, $\tilde{Z}$ has a zero-mean Gaussian distribution with variance $\frac{a^2P+\sigma_E^2}{g^2}$, i.e. $\tilde{Z}\sim\mathcal{N}(0,\frac{a^2P+\sigma_E^2}{g^2})$. Let us define the random variable $E'$ that takes the values $E'_1$, $E'_2$, and $E'_3$, where $E'_1=\{|\tilde{Z}|<l\}$ is the event that the signal before Eve's A/D falls in its dynamic range, and the events $E'_2=\{\tilde{Z}>l\}$ and $E'_3=\{\tilde{Z}<-l\}$ correspond to clipping (A/D overflow). We have, \[ h(Z)=h(Z|E')+H(E')-H(E'|Z), \] Since $E$ is completely determined by $Z$, $H(E|Z)=0$; thus, \begin{align*} h(Z)&=\sum_{i=1}^3h(Z|E'_i)p(E'_i)-\sum_{i=1}^3 p(E'_i)\log (p(E'_i)). \end{align*} In the case of clipping we have $h(Z|E'_2)=h(Z|E'_3)=0$. The probability that the A/D is not in overflow is: \[ p(E'_1)=1-2Q\left(\frac{gl}{\sqrt{a^2P+\sigma_E^2}}\right), \] and the probability that her A/D overflows is given by: \[ p(E'_2)=p(E'_3)=Q\left(\frac{gl}{\sqrt{a^2P+\sigma_E^2}}\right), \] Then, $h(Z|E'_1)$ is calculated as: \begin{align} \nonumber &f_{Z|E'_1}(z)=f_{\tilde{Z}|E'_1}(z)*f_{n_q}(z)\\\nonumber &=\frac{1}{\delta}\int_{-l}^{l}f_{\tilde{Z}}(s)U_{[-{\delta}/{2},{\delta}/{2}]}(z-s)ds\\\nonumber &=\frac{1}{\delta}\int_{\max(-l,z-\delta/2)}^{\min(l,z+\delta/2)}f_{\tilde{Z}}(s)ds\\\nonumber &\approx \frac{1}{\delta}\int_{z-\delta/2}^{z+\delta/2}f_{\tilde{Z}}(s)ds\\ &=\frac{1}{\delta}\left(Q\left(\frac{g(z-\delta/2))}{\sqrt{a^2P+\sigma_E^2}}\right)-Q\left(\frac{g(z+\delta/2)}{\sqrt{a^2P+\sigma_E^2}}\right)\right),\: |z|<l \label{eq:n4} \end{align} where $U_{[-{\delta}/{2},{\delta}/{2}]}(.)$ is the rectangle function on $[-{\delta}/{2},{\delta}/{2}]$, i.e. the value of the function is 1 on the interval $[-{\delta}/{2},{\delta}/{2}]$ and is zero elsewhere. The reason that the approximation is valid is that we assume high precision A/Ds are applied and thus $\delta \ll l$. Hence, \begin{align}\nonumber h(Z)=&\left(1-2Q\left(\frac{gl}{\sqrt{a^2P+\sigma_E^2}}\right)\right)\\ &\int_{-l}^l -f_{Z|E'_1}(z)\log(f_{Z|E'_1}(z)) dz +H(E'). \label{eq:n5} \end{align} Similarly, for $h(Z|X)$ we have, \[h(Z|X)=h(Z|X,E')+H(E'|X)-H(E'|X,Z) \] Since $H(E'|X,Z)=0$, \begin{align} h(Z|X)=\sum_{i=1}^3h(Z|E'_i,X)p(E'_i|X)+H(E'|X) \label{eq:n7} \end{align} where $h(Z|E'_2,X=x)=h(Z|E'_3,X=x)=0$. The probability that Eve's A/D works in its dynamic range given $X$ is, \begin{align*} p(E'_1|X=x)&=p(|\tilde{Z}|<l|X=x)\\ &=p(|\frac{ax}{g}+\frac{n_E}{g}|<l)\\ &=p(-(gl+Ax)<n_E<gl-Ax))\\ &=Q\left(\frac{-(gl+Ax)}{\sigma_E}\right)-Q\left(\frac{gl-Ax}{\sigma_E}\right) \end{align*} and the probability that her A/D overflows, \begin{align*} p(E'_2|X=x)&=p(\tilde{Z}>l|X=x)\\ &=p(\frac{ax}{g}+\frac{n_E}{g}>l)\\ &=Q\left(\frac{gl-Ax}{\sigma_E}\right), \end{align*} and, \begin{align*} p(E'_2|X=x)&=p(\tilde{Z}<-l|X=x)\\ &=p(\frac{ax}{g}+\frac{n_E}{g}<-l)\\ &=Q\left(\frac{gl+Ax}{\sigma_E}\right). \end{align*} In order to calculate $h(Z|E'_1,X)$, $f_{Z|E'_1,X=x}(z)$ is required. The signal before Eve's A/D $\tilde{Z}$ given $X=x$ has a Gaussian distribution with mean $Ax/g$ and variance ${\sigma_E^2}/{g^2}$ within interval $|{ax}/{g}+{n_E}/{g}|<l$ and zero elsewhere. Hence, \begin{align*} &f_{Z|E'_1,X=x}(z)=f_{\tilde{Z}|E'_1,X=x}(z)*f_{n_q}(z)\\ &\approx\frac{1}{\delta}\int_{z-\delta/2}^{z+\delta/2} f_{\tilde{Z}|X=x}(z)ds\\ &=\frac{1}{\delta}\left(Q\left(\frac{g(z-\delta/2)-Ax}{\sigma_E}\right)-Q\left(\frac{g(z+\delta/2)-Ax}{\sigma_E}\right)\right), \end{align*} for $|z|<l$, and, \begin{align} \nonumber &h(Z|X)=\\ \nonumber &\int_{-\infty}^{\infty}\Big[\int_{-l}^l -f_{Z|E'_1,X=x}(z)\log(f_{Z|E'_1,X=x}(z))dz\: p(E'_1|X=x)\\ &\quad -\sum_{i=1}^3p(E'_i|X=x)\log(p(E'_i|X=x))\Big] f_X(x) dx \label{eq:n6} \end{align} By substituting $h(Z)$ from (\ref{eq:n5}) and $h(Z|X)$ from (\ref{eq:n6}) in the following equation, \begin{equation} I(X;Z)=h(Z)-h(Z|X),\label{eq:n13} \end{equation} the mutual information between Alice and Eve given $A=a$ and $G=g$, can be found. Also, by substituting $Z$ with $Y$, $\sigma_E^2$ with $\sigma_B^2$, and $g$ with $a$ in (\ref{eq:n5}), (\ref{eq:n6}), and (\ref{eq:n13}), the mutual information between Alice and Bob given $A=a$ can be found, \begin{equation} I(X;Y)=h(Y)-h(Y|X)\label{eq:n12} \end{equation} The achievable secrecy rate can be found by substituting these mutual informations into the following equation: \begin{align} R_s=E_{G,A}\left[I(X;Y)-I(X;Z)\right] \label{eq:9} \end{align} Alice is able to choose $p$ to maximize the $R_s$ that can be achieved by this method; on the other side, Eve tries to minimize $R_s$ by choosing an appropriate $p_G(.)$. The following lemma shows that for an arbitrary discrete alphabet for $G$, choosing a single value (which depends on the value of $p$) with probability one minimizes the secrecy rate, and thus is the optimal strategy for Eve. \begin{lem} The gain $1/G$ that Eve applies before her A/D should take a single value with probability one to minimize the secrecy rate. \end{lem} \begin{proof} Suppose $G$ has the following probability mass function: \[ p_G(g=G_i)=\alpha_i, \quad \quad i=1,\cdots,n\\ \] such that $\sum_{i=1}^{n}\alpha_i =1$. Without loss of generality, assume that for a specific $p$, the maximum information leakage occurs at $G=G_1$, i.e. for any gain $G_i, i=2,\cdots,n$ we have $I(X;Z|G=G_1)\geq I(X;Z|G=G_i)$; hence, \begin{align*} I(X;Z)&=\sum_{i=1}^{n}\alpha_i I(X;Z|G=G_i)\\ & \leq \sum_{i=1}^{n}\alpha_i I(X;Z|G=G_1)=I(X;Z|G=G_1) \end{align*} \end{proof} The above lemma can easily be generalized to continuous random variables. Numerical results are given in Sections \ref{sec:4} and \ref{sec:5}. \subsection{Noiseless Channels} In the case that the channel between Alice and Eve is noiseless, $h(Z)$ can be found by setting $\sigma_E^2=0$ in (\ref{eq:n5}). Using (\ref{eq:n6}) and the fact that $h(Z|E'_2,X=x)=h(Z|E'_3,X=x)=0$ and $H(E'|X)=0$ we have, \begin{align} \nonumber &h(Z|X)\\ \nonumber&=\int_{-\infty}^{\infty}h(Z|E'_1,X=x)p(E'_1|X=x)f_X(x)dx \\ \nonumber&=\int_{-\infty}^{\infty}h(\frac{aX}{g}+n_q|E'_1,X=x)p(E'_1|X=x)f_X(x)dx\\ \nonumber&=\int_{-Gl/|A|}^{gl/|A|}h(n_q)f_X(x)dx\\ &=\log(\delta)\left(1-2Q\left(\frac{gl}{a\sqrt{P}}\right)\right) \label{eq:n10} \end{align} Similarly, in the case that Bob has a noiseless channel, \begin{equation} h(Y|X)=\log(\delta)\left(1-2Q\left(\frac{l}{\sqrt{P}}\right)\right)\label{eq:n11} \end{equation} In each case, the secrecy rate can be found by substituting (\ref{eq:n10}) and (\ref{eq:n11}) in (\ref{eq:n13}) and (\ref{eq:n12}), respectively. Numerical results for the noiseless channels are shown in Sections \ref{sec:4} and \ref{sec:5}. \begin{figure} \begin{center} \includegraphics[width=.4\textwidth]{wiretap_channel_erasure3.eps} \end{center} \caption{Gaussian erasure wiretap channel: in the asymptotic case, the erasures/overflows at Eve's A/D due to the rapid power modulation at the transmitter can be modeled by an erasure channel. } \label{fig:pic5} \end{figure} Clearly, considering noiseless channels makes the results less complicated and thus more insightful. Hence, we continue our investigation by studying the asymptotic behavior of the proposed method (as $r\rightarrow \infty$) in the noiseless regime, which will help us to achieve some intuition regarding this scheme. We assume that Bob and Eve use A/D's of the same quality for this analysis. Since in the noiseless regime $I(X;Y)$ does not depend on $A$, it does not change with $r$ and thus we need only evaluate $I(X;Z)$ for our asymptotic analysis. From (\ref{eq:cons}) we have, \begin{equation}\label{eq:As} A_1=\frac{r}{\sqrt{pr^2+(1-p)}}\quad\text{and}\quad A_2=\frac{1}{\sqrt{pr^2+(1-p)}} \end{equation} Let $G(r)$ be the inverse of the gain that Eve employs as a function of $r$. Recall from Lemma 1 that $G(r)$ will take a single value with probability one for a given $r$, but that value can depend on $r$. Since $A_1\rightarrow 1/\sqrt{p}$ and $A_2\rightarrow 0$, we claim that in the limit (as $r\rightarrow \infty$), the best strategy that Eve can take is to choose either $G(r)=\Theta(1)$ or $G(r)=\Theta(r^{-1})$; otherwise, she will get no information (see Appendix A). First we study the secrecy rates that can be achieved when $G(r)=\Theta(1)$ as $r$ approaches $\infty$. The average secrecy rate is: \begin{align} \nonumber R_s&=E[I(X;Y)-I(X;Z)]\\ \nonumber &=p(I(X;Y|A=A_1)-I(X;Z|A=A_1))\\ &\quad +(1-p)(I(X;Y|A=A_2)-I(X;Z|A=A_2))\label{eq:ER} \end{align} Assuming that Bob chooses the optimum range for his A/D, the maximum $I(X;Z|A=A_1)$ that Eve can achieve is $I(X;Y|A=A_1)$ and hence the first term in (\ref{eq:ER}) is zero. To evaluate the second term, putting $G(r)$ and $A=A_2$ in (\ref{eq:n5}) and (\ref{eq:n10}) yields: \begin{align}\nonumber h(Z)=&\left(1-2Q\left(\frac{G(r)l}{A_2\sqrt{P}}\right)\right)\\ &\int_{-a}^a -f_{Z|E'_1}(z)\log(f_{Z|E'_1}(z)) dz+H(E') \label{eq:ne13} \end{align} where since $G(r)=\Theta(1)$, $\left(1-2Q\left(\frac{G(r)l}{A_2\sqrt{P}}\right)\right)\to 1$ as $r\to \infty$ and thus $H(E')\to 0$; and, for $|z|<l$, \begin{align*} &f_{Z|E'_1}(z)\\ &=\frac{1}{\delta}\left(Q\left(\frac{G(r)(z-\delta/2)}{A_2\sqrt{P}}\right)-Q\left(\frac{G(r)(z+\delta/2)}{A_2\sqrt{P}}\right)\right) \\ &\rightarrow \left\{ \begin{array}{l l} \frac{1}{\delta}, & \quad 0<|z|<\delta/2\\ \frac{1}{2\delta}, & \quad |z|=\delta/2\\ 0, & \quad \quad \text{otherwise}\\ \end{array} \right. \end{align*} Since the integrand in (\ref{eq:ne13}) is bounded for all $r$, from the dominated convergence theorem, $h(Z)\rightarrow \log\delta$ as $r\rightarrow \infty$. Also since $G(r)=\Theta(1)$, \begin{align} h(Z|X)&=\log(\delta)\left(1-2Q\left(\frac{G(r)l}{A_2\sqrt{P}}\right)\right) \rightarrow \log(\delta) \label{eq:n14} \end{align} as $r$ approaches $\infty$. Thus, $I(X;Z|A=A_2)=0$ and hence the average secrecy rate given that $G(r)=\Theta(1)$ is $R_s=(1-p)I(X;Y)$. Now suppose $G(r)=\Theta(r^{-1})$ and consider the second term in (\ref{eq:ER}). In the limit, ${A_2}/{G(r)}=c$ where $c>0$ is a bounded constant. Since Bob chooses the optimum range for his A/D, the maximum $I(X;Z|A=A_2)$ that Eve can achieve is $I(X;Y|A=A_2)$ and thus given that $G(r)=\Theta(r^{-1})$, the second term in (\ref{eq:ER}) is zero. To evaluate the first term in (\ref{eq:ER}) as $r$ gets large, by substituting $G(r)=\Theta(r^{-1})$ and $A=A_1$ in (\ref{eq:n5}) and (\ref{eq:n10}), we have $f_{Z|E'_1}(z)\rightarrow 0$ and, \[ \left(1-2Q\left(\frac{G(r)l}{A_1\sqrt{P}}\right)\right)\rightarrow 0 \] as $r$ approaches infinity and hence $h(Z)\rightarrow 0$. Also by letting $G(r)=\Theta(r^{-1})$ we have, \begin{align*} h(Z|X)&=\log(\delta)\left(1-2Q\left(\frac{G(r)l}{A_1\sqrt{P}}\right)\right)\rightarrow 0 \quad\text{as} \quad r\rightarrow\infty \end{align*} Hence, with probability $p$ the mutual information between Alice and Eve is zero and the average secrecy rate that can be achieved given $G(r)=\Theta(r^{-1})$ as $r$ approaches $\infty$ is $R_s=pI(X;Y)$. We can interpret these results as follows; when ${A}/{G(r)}={A_1}/{\Theta(r^{-1})}$, the total gain that Eve's A/D sees approaches infinity as $r \rightarrow \infty$; hence, even if Eve uses an A/D with larger range than Bob's A/D, her quantizer overflows. When ${A}/{G(r)}={A_2}/{\Theta(1)}$, the total gain goes to zero as $r$ approaches infinity and thus even if Eve uses an A/D with better precision, the received signal amplitude is less than one quantization level. In both cases, Eve receives no information about the transmitted signal and thus Eve's channel can be modeled by an erasure channel (Figure \ref{fig:pic5}), where for $G(r)=\Theta(r^{-1})$, the probability of erasure $\epsilon=1-p$ and for $G(r)=\Theta(1)$, $\epsilon=p$. Hence, the secrecy rate that can be achieved in the asymptotic case (as $r\rightarrow \infty$) is: \begin{equation} R_s=(1-\epsilon) I(X;Y) \label{eq:Rs} \end{equation} \begin{figure} \begin{center} \includegraphics[width=.5\textwidth]{versus_r.eps} \end{center} \caption{Achievable secrecy rate versus $r$ (the ratio between the large and the small gain); both Bob and Eve apply 10-bit A/D's with the dynamic range $l=2.5$. The SNRs of both Bob's channel and Eve's channel are the same and are denoted by $\alpha$.} \label{fig:f7} \end{figure} To maximize the achievable secrecy rate, it is reasonable for Alice to choose $p=0.5$. In Section \ref{sec:s1} it is shown that for a 10-bit A/D and the transmitter power $P=1$, the optimum range of the A/D is obtained by setting $l=2.5$, and the corresponding mutual information between Alice and Bob (when the channel between them is noiseless) is $I(X;Y)=6.597$. Hence, using (\ref{eq:Rs}), $R_s\rightarrow 0.5\times 6.597=3.2985$. Figure \ref{fig:f7} (the upper curve) shows the achievable secrecy rate versus $r$ when both Bob's channel and Eve's channel are noiseless. It can be seen that as $r$ gets larger, the achievable secrecy rate goes to a constant which is similar to what is anticipated. Furthermore, for larger $r$'s ($r\geq10^3$) the optimum probability that maximizes the worst case secrecy rate is $p=0.5$. These results show that our results are consistent to expectations in the limit. From another point of view, consider that for small values of $\delta$, the quantization noise can be modeled by a zero mean Gaussian random variable with the variance ${\delta^2}/{12}$, where $\delta$ is the size of each quantization level. Thus, this wiretap channel can be modeled by a Gaussian erasure wiretap channel. The secrecy capacity of the Gaussian wiretap channel is \cite{barros2006secrecy}: \[C_s=\frac{1}{2}\left(\log(1+|h_B|^2\gamma_B)-\log(1+|h_E|^2\gamma_E)\right)^+\] where $h_B$ and $h_E$ are channel gains, $\gamma_B$ is the SNR at Bob's receiver, and $\gamma_E$ is the SNR at Eve's receiver. We can use this secrecy capacity in our asymptotic model by setting $h_B=1$ and modeling the erasure channel by an unusual fading channel with the following fading distribution: \[h_E=\left\{ \begin{array}{l l} 0,&\quad \text{w.p.}\quad \epsilon\\ 1,&\quad \text{w.p.}\quad 1-\epsilon\\ \end{array} \right. \] Since we assumed that Eve's A/D is identical to Bob's A/D, $\gamma_E=\gamma_B=\frac{P}{\nicefrac{\delta^2}{12}}$ and thus the secrecy capacity is non-zero only when an erasure at Eve's channel occurs. Hence, \begin{equation} C_s=\frac{(1-\epsilon)}{2} \log(1+\gamma_B) \label{eq:Cs} \end{equation} This equation shows that for a 10-bit A/D with $l=2.5$, transmitting power $P=1$, and $\epsilon=p=0.5$, the secrecy capacity is $C_s=3.2822$ which is again very close to what we expect from our asymptotic analysis. Furthermore, on comparing equations (\ref{eq:Rs}) and (\ref{eq:Cs}), it is seen that in the asymptotic case, the achievable secrecy rate meets this approximate secrecy capacity. \section{Numerical Results} \label{sec:4} \subsection{Motivation}\label{sec:41} \begin{figure} \begin{center} \includegraphics[width=.5\textwidth]{PD.eps} \end{center} \caption{Secrecy capacity of public discussion for various values of SNR at Bob's receiver when the SNR at Eve's receiver changes from 50 dB to 100 dB and $P=1$. When the SNR at Bob's receiver is less than the SNR Eve's receiver, the secrecy rate drops rapidly.} \label{fig:p5} \end{figure} When the channel between Alice and Eve is less noisy than the channel between Alice and Bob, if the legitimate users are restricted to one-way and rate-limited communication, the secrecy capacity of the wiretap channel is zero. However, if we relax the restrictions placed on the schemes that the legitimate users can apply by allowing two-way communication and the presence of a noiseless, public, and authenticated channel, public discussion strategies \cite{maurer1993secret,ahlswede1993common} allow the legitimate nodes to agree on a secret key by extracting information from realizations of correlated random variables. This secret-key can then be used in a one-time-pad for secret communication between Alice and Bob. A closed form for the general secret-key capacity is not available; however, in the case of a Gaussian source model in which $X\sim\mathcal{N}(0,P)$ and a Gaussian wiretap channel, i.e. when the channel between Alice and Bob and the channel between Alice and Eve are AWGN channels, the secrecy capacity has a simple form \cite[Chapter 5]{bloch2011physical}: \begin{equation} C^{\text{SM}}_s=\frac{1}{2}\log\left(1+\frac{P\sigma_E^2}{(P+\sigma_E^2)\sigma_B^2}\right) \end{equation} and thus all secret-key rates less than $C^{\text{SM}}_s$ are achievable. Achievable secrecy rates of public discussion for various values of the signal-to-noise ratio at Bob's receiver versus signal-to-noise ratio at Eve's receiver are shown in Figure \ref{fig:p5}. As can be seen, when the SNR of Eve's receiver is significantly larger than the SNR at Bob's receiver, the secrecy rate of public discussion drops rapidly. Our main goal here is to see whether our scheme can improve the performance in this regime. \subsection{Noiseless Channels: Eve with the same A/D as Bob}\label{sec:s1} \begin{figure} \begin{center} \includegraphics[width=.39\textwidth]{Noiselesstransparent.eps} \end{center} \caption{Achievable secrecy rate vs. the probability $p$ and the gain $G$ at Eve's receiver. Both Bob's and Eve's channels are noiseless and they use identical 10-bit A/D's. The ratio between the two power levels at the transmitter is $r=10^3$ (i.e. 30 dB) and the average transmitting power is $P=1$. A maxi-min rate of $R_s=3.1372$ is achieved.} \label{fig:p2} \end{figure} We begin our investigation by considering only the effect of A/D's on the signals. Hence, we assume that Eve's channel is noiseless, i.e. $n_E=0$ (which benefits the eavesdropper). However, we also assume the system nodes are working in a very high SNR regime and thus the channel noise at Bob can be neglected ($n_B=0$). Now suppose that both Bob and Eve use 10-bit quantizers ($b=2^{10}$) and the transmitter power is $P=1$. Since $\delta=2a/b$, for a fixed number of quantization bits, $I(X;Y)$ is a function of the of the A/D ($a$), and the optimal quantization range that maximizes $I(X;Y)$ can be found. Since $I(X;Y)$ is an intricate function in terms of $a$, we find the optimum $a$ numerically. In this case, the optimum quantization range that maximizes $I(X;Y)$ is $l=2.5$, and the corresponding mutual information between Alice and Bob is $I(X;Y)=6.597$. For the remainder of the paper, we use $l=2.5$ in our calculations. Suppose that Eve has the same A/D as Bob. From Lemma 1, putting a random gain is undesirable for Eve; hence, she chooses a fixed gain $G$ that minimizes $R_s$. Because Alice is not aware of Eve's choice, she has to choose a probability $p$ that maximizes the worst case $R_s$. As we discussed in Section \ref{sec:2}, a larger $r$ leads to more eavesdropper confusion and thus as $r$ increases, the secrecy rate would be expected to increase. However, in the case of noisy channels, a large $r$ also causes noise enhancement at Bob's receiver that decreases the secrecy rate. In order to get some insight about the dependency of the secrecy rate on $r$, curves of $R_s$ versus $r$ are shown in Figure \ref{fig:f7}. For each curve, the SNR at both Eve's receiver and Bob's receiver are the same and are denoted by $\alpha$. Hence, in order to achieve high secrecy rates by avoiding excessive noise enhancement at Bob's receiver, for the rest of the paper we set $r=10^3$. The plot of $R_s$ versus $p$ and $G$ for $P=1$ and $r=10^3$ (i.e. 30 dB) where both Bob and Eve are each using a 10-bit A/D is shown in Figure \ref{fig:p2}. This function is complicated and hence the optimum value of $p$ cannot be derived analytically. Numerical analysis shows that $p\approx 0.45$ maximizes the worst case $R_s$, and the maxi-min value is ${R_s}=3.1366$. Hence, choosing $p=0.45$ guarantees that at least the secrecy rate ${R_s}=3.1366$ can be achieved. \subsection{Noiseless Channels: Eve with a Better A/D than Bob} \begin{figure} \begin{center} \includegraphics[width=.39\textwidth]{14bitADvs10bitAD.eps} \end{center} \caption{Achievable secrecy rate vs. the probability $p$ and the gain at Eve's receiver, $G$ for the case of noiseless channels. The ratio between the two power levels at the transmitter is $r=10^3$ (i.e. 30 dB) and the average transmitting power is $P=1$. In the upper curve, both Bob and Eve have the same 10-bit A/D's. In the lower curve, Bob uses a 10-bit A/D while Eve uses a 14-bit A/D (Eve's A/D is 24 dB better than Bob's A/D) and a maxi-min rate of $R_s=1.2478$ is achieved (for $p=0.4$). } \label{fig:p3} \end{figure} Now suppose that Eve has access to a better A/D than Bob. Depending on the gain that Eve applies before her A/D, a better A/D results in less erasures and/or less A/D overflows. Hence, the mutual information between Alice and Eve increases and consequently, the achievable secrecy rate decreases. Figure \ref{fig:p3} shows this effect versus $p$ and $G$. It can be seen that even if Eve uses an A/D which is 24 dB (4 bits) better than Bob's A/D (Eve has a 14-bit A/D while Bob has a 10-bit A/D), by choosing an appropriate value for $p$, a positive secrecy rate can be achieved. In this example, by choosing $p=0.4$, a secrecy rate $R_s=1.2426$ is achievable. Even if we do not change the probability $p$ from the previous section ($p=0.45$), assuming that Alice is not aware of Eve's better A/D, a secrecy rate $R_s=0.9225$ is achievable. In spite of having a better A/D, Eve will still lose some symbols and hence a positive secrecy rate is available. This is because the ratio between the large and the small gain, $A_1$ and $A_2$, is $10^3$, while Eve's A/D has only 16 times better resolution; thus, she still needs to compromise between resolution and overflow. To cancel the effect of these gains completely, Eve has to use an A/D that has an effective resolution after taking into account jamming, interference, etc. on the order of $10^3$ times (10 bits) better than Bob's A/D, which would be very difficult in an adversarial environment. \subsection{Noisy Main Channel, Noiseless Eavesdropper's channel}\label{sec:s3} \begin{figure} \begin{center} \includegraphics[width=.5\textwidth]{NoisyBobNoiselessEve.eps} \end{center} \caption{Achievable secrecy rate vs. SNR at Bob's receiver while the SNR at Eve's receiver is infinity (Eve has perfect access to the transmitted signal) for $r=10^3$, $P=1$ and Bob and Eve applying 10-bit A/D's. Note that the assumption of Eve having a noiseless channel is the extreme case when Eve has perfect access to the transmitter's output (for instance, the eavesdropper is able to pick up the transmitter's radio) and hence other secrecy methods are not effective. Using the proposed method, a positive secrecy rate can be achieved over short range at reasonable power.} \label{fig:p4} \end{figure} Now we look at the extreme case that Eve is able to receive exactly what Alice transmits and receives (e.g. the adversary is able to pick up the transmitter's radio and hook directly to the antenna), but the channel between Alice and Bob is noisy and hence no other technique is effective. In other words, the channel between Alice and Bob experiences an additive white Gaussian noise ($n_B\sim\mathcal{N}(0,\sigma_B^2)$), while Eve's channel is noiseless ($n_E=0$). Figure \ref{fig:p4} shows the secrecy rate $R_s$ that can be achieved using the proposed scheme versus the signal-to-noise ratio (SNR) at Bob's receiver. In this case, the transmitted power $P=1$, the ratio between the large and the small gain is 30 dB, and both Bob and Eve use 10-bit A/D's. It can be seen that, although Eve's channel is much better than Bob's channel, when the SNR at Bob's receiver is greater than 60 dB, which could be made common in a short-range application as described in Section \ref{sec:1}, a positive secrecy rate is available. By comparing the noise-free result in Figure \ref{fig:f7} for $r=10^3$ and Figure \ref{fig:p4}, it can be seen that the secrecy rate when SNR at Bob is 120 dB is still less than the secrecy rate when Bob's channel is noiseless. \subsection{Noisy Channels}\label{sec:5} When both channels are noisy, the achievable secrecy rate of the proposed method versus the SNR at Eve's receiver for various values of the SNR at Bob's receiver is shown in Figures \ref{fig:p6}. The transmitted power $P=1$, the ratio between the large and the small gain is 30 dB, and both Bob and Eve use 10-bit A/D's. It can be seen that by applying the proposed method for the case of Eve with a (significantly) better channel than Bob, which is the regime of interest per Figure \ref{fig:p5}, reasonable secrecy rates can be achieved. Note that in our method we are generating an advantage for the legitimate nodes to be used with wiretap coding, and thus, because public discussion approaches assume the presence of a public authenticated channel, public discussion should not be viewed as a competitor to the proposed scheme. Rather, if such a public authenticated channel exists and two-way communication is possible, our method can be used \textit{in conjunction} with public discussion techniques to result in higher secrecy rates. Nevertheless, per Figure \ref{fig:p5}, public discussion provides motivation for the regime where advances are needed given the current state of the art. \begin{figure} \begin{center} \includegraphics[width=.5\textwidth]{our_method.eps} \end{center} \caption{Achievable secrecy rates for various values of the SNR at Bob's receiver when the SNR at Eve's receiver changes from 50 dB to 100 dB. The settings are $r=10^3$, $P=1$, and Bob and Eve are applying 10-bit A/D's. When the SNR of the channel between Alice and Eve is significantly better than the SNR of the channel between Alice and Bob, reasonable secrecy rates are still achievable.} \label{fig:p6} \end{figure} \section{Conclusion}\label{sec:7} In this paper, we introduce a new approach that exploits a short-term cryptographic key to force different orderings at Bob and Eve of two operators, one of which is necessarily non-linear, to obtain the desired advantage for information-theoretic security in a wireless communication system regardless of the location of Eve. We then investigate a simple power modulation instantiation of the approach. It is shown that when Eve's channel condition is significantly better than the Bob's channel, reasonable secrecy rates can still be achieved using our proposed method in this challenging regime. In particular, even in the case that the adversary is able to pick up the transmitter's radio (i.e. Eve has perfect access to the output of the transmitter), a reasonable secrecy rate is achievable at high SNRs which might apply to a short-range wireless system. For example, one might use the transmission power of typical cellular systems with the corresponding excess power at short ranges to establish a secure radio system in a limited area. Although we have considered the case of Eve with a better A/D than Bob, the clear risk to the approach is still that of asymmetric capabilities at the receivers. For example, if we employ the simple power modulation approach studied extensively here, Eve may employ multiple A/D's with different gain settings in front of each. Hence, Eve would be able to record two signals independently and decode them later when she gets the key or extracts the key based on the pattern of erasures and overflows at each A/D. A simple approach to combat this attack is rather than applying just two power levels, the transmitter can apply many power levels. More promising, however, is to consider adding memory to the signal warping process \cite{isit2013}. Broadly considering potential techniques for everlasting security in wireless systems, including that proposed here, yields that each approach still holds some risk. In the case of cryptographic security, assumptions must be made on both the hardness of the problem and the current/future computational capabilities of the adversary. In the case of standard information-theoretic security, assumptions must be made on the quality of the channel to Eve, generally corresponding to limitations on her location. In the method proposed here, assumptions must be made on Eve's current conversion hardware capabilities, but, as in standard information-theoretic secrecy, there is no assumption on future capabilities. All three approaches thus have different applicability. \section*{Appendix A} In this section we show that as $r\to\infty$ the only strategy that Eve can take to obtain information from the signal she receives is to choose either $G(r)=\Theta(1)$ or $G(r)=\Theta(r^{-1})$. Instead of applying $G(r)=\Theta(1)$ or $G(r)=\Theta(r^{-1})$, the two other possibilities for Eve are to choose $G(r)$ such that either $\lim_{r \to \infty} r^{-1}/G(r)\to 0$ or $\lim_{r \to \infty} r^{-1}/G(r)\to \infty$ (and obviously provided that $G(r)\neq\Theta(1)$). First suppose $\lim_{r \to \infty} r^{-1}/G(r)\to 0$ and consider $I(X;Z|A=A_1)$ in (\ref{eq:ER}). Since $G(r)\neq\Theta(1)$ and from (\ref{eq:As}), $\lim_{r \to \infty} A_1/G(r)\to 0$ and hence, \begin{align} h(Z)=\left(1-2Q\left(\frac{G(r)l}{A_1\sqrt{P}}\right)\right)\int_{-a}^a -f_{Z|E'_1}(z)\log(f_{Z|E'_1}(z)) dz \label{eq:nee13} \end{align} where, for $|z|<l$, \begin{align*} f_{Z|E'_1}(z)&=\frac{1}{\delta}\left(Q\left(\frac{G(r)(z-\delta/2)}{A_1\sqrt{P}}\right)-Q\left(\frac{G(r)(z+\delta/2)}{A_1\sqrt{P}}\right)\right) \\ &\rightarrow \left\{ \begin{array}{l l} \frac{1}{\delta}, & \quad 0<|z|<\delta/2\\ \frac{1}{2\delta}, & \quad |z|=\delta/2\\ 0, & \quad \quad \text{otherwise}\\ \end{array} \right. \end{align*} and $\left(1-2Q\left(\frac{G(r)l}{A_1\sqrt{P}}\right)\right)\rightarrow 1$ as $r\rightarrow \infty$. Since the integrand in (\ref{eq:nee13}) is bounded for all $r$ and from the dominated convergence theorem, $h(Z)\rightarrow \log\delta$ as $r\rightarrow \infty$. Also, since $\lim_{r \to \infty} r^{-1}/G(r)\to 0$, \begin{align} h(Z|X)&=\log(\delta)\left(1-2Q\left(\frac{G(r)l}{A_1\sqrt{P}}\right)\right) \rightarrow \log(\delta) \label{eq:500} \end{align} as $r$ approaches $\infty$ and thus $I(X;Z|A=A_1)=0$. Now consider $I(X;Z|A=A_2)$ in (\ref{eq:ER}); by substituting $A_1$ with $A_2$ in (\ref{eq:nee13}) and (\ref{eq:500}), and since $\lim_{r \to \infty} A_2/G(r)\to 0$, we have $I(X;Z|A=A_2)=0$. Consequently, given that $\lim_{r \to \infty} r^{-1}/G(r)\to 0$, the average information that Eve obtains is zero. Now suppose $\lim_{r \to \infty} r^{-1}/G(r)\to \infty$ and consider the first term $I(X;Z|A=A_1)$ in (\ref{eq:ER}). The fact that $\lim_{r \to \infty} r^{-1}/G(r)\to \infty$ implies that in the limit as $r\to\infty$, $ A_1/G(r)$ also goes to $\infty$ and thus from (\ref{eq:n5}) and (\ref{eq:n10}) we have $f_{Z|E'_1}(z)\rightarrow 0$. Also, $\left(1-2Q\left(\frac{G(r)l}{A_1\sqrt{P}}\right)\right)\rightarrow 0$ as $r$ approaches infinity and hence $h(Z)\rightarrow 0$. Furthermore, \begin{align}\label{eq:hzx} h(Z|X)&=\log(\delta)\left(1-2Q\left(\frac{G(r)l}{A_1\sqrt{P}}\right)\right)\rightarrow 0 \end{align} as $r\to\infty$ and thus $I(X;Z|A=A_1)=0$. Considering $I(X;Z|A=A_2)$ in (\ref{eq:ER}) and by putting $A_2$ instead of $A_1$ in (\ref{eq:hzx}), since $A_2/G(r)\rightarrow \infty$ in the limit as $r\rightarrow \infty$, we have $I(X;Z|A=A_2)=0$. Hence, by choosing $\lim_{r \to \infty} r^{-1}/G(r)\to \infty$ Eve gets no information about the transmitted signal. \bibliographystyle{ieeetr}
\section{Introduction} \hspace{.25in} The well-known hyperbolic, or Poincar\'{e}, density on plane domains has proved to be of great importance and utility in complex analysis and geometry. It is known that if $U$ be a hyperbolic domain and $U_n$ a sequence of domains in $\mathbb{C}$ which converge to $U$ in a reasonable way, then the hyperbolic density on $U_n$ converges locally uniformly to that on $U$. However, for applications it may be useful to be able to say something about the rate of this convergence. This is the question, originally posed by A. Douady, to which this paper is devoted. \hspace{.25in} Below, all distances are measured to the Euclidean metric unless otherwise specified. The Hausdorff distance $H(A,B)$ between any two sets $A$ and $B$ is defined to be \begin{equation} \label{} H(A,B):= \inf_{r>0} \{A \subseteq N_r(B) \mbox{ and } B \subseteq N_r(A) \}, \end{equation} where $N_r(A)$ is the set of all points whose distance from $A$ is less than $r$. In contrast, we define \begin{equation} \label{} d(A,B):= \inf_{z\in A,w \in B} \{|z-w|\}, \end{equation} For example, if we let $A=\{1/2\}, B = \{|z|=1\}$, and $C=\{|z-2|=1\}$, then $H(A,B)=3/2, H(A,C)=5/2, H(B,C)=2, d(A,B)=1/2, d(A,C)=1/2, d(B,C)=0$. For singleton sets we will commonly use the shorthands $H(x,A), d(x,A)$ instead of $H(\{x\},A), d(\{x\},A)$. The relation between $d$ and $H$ is that \begin{equation} \label{} H(A,B) = \max\Big(\sup_{z \in A} d(z,B) , \sup_{z \in B} d(z,A) \Big) \end{equation} The following is the definition of convergence of domains which will be useful to us. \begin{deff} \label{d1} Let $U$ be a domain and $U_n$ a sequence of domains in $\mathbb{C}$. We say that $U_n$ converges in boundary to $U$ if \begin{itemize} \item[a)]{$H(\delta U, \delta U_n) \longrightarrow 0$} \item[b)]{There exists $z_0 \in U$ such that $z_0 \in U_n$ for all $n$.} \end{itemize} \end{deff} {\bf Remark:} Condition $(b)$ is necessary to eliminate such situations as $U=\mathbb{D}$, $U_n = \{1-1/n < |z| < 1\}$. \vspace{12pt} A related concept is {\it Carath\'{e}odory convergence}. Consider a sequence of domains $\{U_n\}$ containing the fixed point $z_0$. The {\it kernel} $U$ of this sequence is defined to be the maximal domain containing $z_0$ such that every compact $K \subseteq U$ is contained in $U_n$ for sufficiently large $n$. If no such $U$ exists, then let $U = \{z_0\}$. We say that $U_n$ converges to its kernel $U$ in the Carath\'{e}odory sense if $U$ is also the kernel of every subsequence of $\{U_n\}$ (see \cite{dur} or any of a number of other books on univalent functions or geometric function theory). The following lemma shows the connection between the two modes of convergence. \begin{lemma} If $U_n \longrightarrow U$ in boundary then $U_n \longrightarrow U$ in the Carath\'{e}odory sense. \end{lemma} {\bf Proof:} Let $z_0$ be as in Definition \ref{d1}, and suppose $K \subseteq U$ is compact. We may assume $K$ is connected and contains $z_0$, as it will always be contained in a compact, connected set within $U$ containing $z$. By compactness, $K$ is a positive distance from $\delta U$. Condition $(a)$ then implies that, for sufficiently large $n$, $K \bigcap \delta U_n = \emptyset$, and thus $K$ lies in $(\delta U_n)^c$. Since $K$ is connected, it lies within a connected component of $(\delta U_n)^c$. As $z \in K$ and $z \in U_n$ for large $n$, it follows that the connected component in question is $U_n$ itself. This shows that the kernel of $\{U_n\}$ is contained in $U$. However, convergence in boundary easily implies \begin{equation} \label{} \bigcap_{N=1}^\infty \bigcup_{n \geq N} U_n \subseteq \bar{U} . \end{equation} Thus, $U$ is the kernel of $\{U_n\}$. It is clear that any subsequence of $\{U_n\}$ also converges in boundary to $U$, and thus by the same argument has $U$ as its kernel. We see that $U_n \longrightarrow U$ in the Carath\'{e}odory sense. {\hfill $\Box$ \bigskip} The converse to this lemma is false, as the domains $U_n=\mathbb{D} \bigcup \{z \in \mathbb{C}: \arg(z) \in (0,\frac{1}{n})\}$ converge to $U=\mathbb{D}$ in the Carath\'{e}odory sense but not in boundary. Since we are interested in the behavior of the hyperbolic density we assume henceforth that $U$ has at least 3 boundary points in the Riemann sphere, which implies that the hyperbolic density for $U$ exists. This clearly implies that if $U_n \longrightarrow U$ in boundary, then $U_n$ is hyperbolic for sufficiently large $n$. Let $\rho_A(z)$ denote the hyperbolic density of any hyperbolic domain $A$ at the point $z$, normalized to have curvature -4 (a normalization with curvature -1 is also common). Then $\rho_A(z) = \frac{1}{|\pi '(0)|}$, where $\pi$ is a holomorphic covering map from $\mathbb{D}$ to $A$ with $\pi(0)=z$. The following lemma should be considered known, though it may not yet have been stated in this form. \begin{lemma} \label{old} If $U_n \longrightarrow U$ in boundary, then $\rho_{U_n}(z) \longrightarrow \rho_U (z)$ locally uniformly. \end{lemma} {\bf Proof:} This is immediate from Theorem 1 of \cite{hej}, which shows under the more general condition of Carath\'{e}odory convergence that the covering maps from $\mathbb{D}$ to $U_n$ converge locally uniformly to that of $U$. {\hfill $\Box$ \bigskip} {\bf Remark:} For a given $z$, $\rho_{U_n}(z)$ may only be defined for large $n$, and the statement of the theorem should be interpreted accordingly. \section{The Teichm\"{u}ller density} \label{tm} \hspace{.25in} In order to say something about the rate of convergence of $\rho_{U_n}(z)$ to $\rho_{U}(z)$, we introduce a different conformally invariant density. Denote the Teichm\"{u}ller density on a domain $A$ by $\lambda_A (z)$, defined as \begin{equation} \label{} \lambda_A (z) := \inf_{V \in {\cal T}} ||\bar{\delta} V||_\infty \end{equation} where $||\cdot||_\infty$ denotes the $L^\infty$ norm, $\bar{\delta}V = \frac{dV}{d\bar{w}} = \frac{1}{2}(\frac{dV}{dx} + i\frac{dV}{dy})$, and ${\cal T}$ is the family of all complex-valued functions $V(w)$ on $A$ with $V(z) = 1$, which vanish on $\delta A$, and with distributional derivatives. Since $\bar{\delta}V$ is bounded, the integral $\int_A \bar{\delta}V \varphi$ converges for every integrable function $\varphi$ on $A.$ By normalizing, we may temporarily assume that 0 and 1 are in the complement of $A.$ Cauchy's formula shows $V(z)=\int_A \bar{\delta}V \varphi_z$ where $\varphi_z(w)=-\frac{z(z-1)}{\pi w(w-1)(w-z)}.$ The linear span of functions of the form $\varphi_z$ where $z$ is in the complement of $A - \{p\}$ is dense in the space of all integrable and holomorphic functions on $A - \{p\}.$ Thus, the chain rule applied to $W(w)=V(f(w))\frac{f'(z)}{f'(w)}$ shows that $\lambda$ is conformally invariant, in the sense that if $f$ is a conformal map we have $\lambda_A(z)=\lambda_{f(A)}(f(z))|f'(z)|$. The Teichm\"{u}ller density has still at this point a relatively brief history. It was originally defined in \cite{fnart}, where it was used in the study of uniformly thick and uniformly perfect domains. It was also proved there that the Teichm\"{u}ller and hyperbolic densities are equivalent, in the sense that $\frac{1}{2}\rho_A \leq \lambda_A \leq \rho_A$ for any domain $A$. The transitivity of the automorphism group of $\mathbb{D}$ together with the conformal invariance of $\lambda_A$ shows that $\lambda_A$ and $\rho_A$ coincide, up to a multiplicative constant, on simply connected domains. It is possible to calculate this constant, and it turns out that $\frac{1}{2}\rho_A = \lambda_A$ in this case. On the other hand, if $A$ is the thrice punctured sphere then it is known that $\rho_A = \lambda_A$ (see \cite{egn}). It was shown in \cite{yu} that $\lambda_A$ is continuous for any domain $A$, and is the infinitesimal form of a previously known metric defined in terms of Teichm\"{u}ller shift mappings(see \cite{kra}). \hspace{.25in} We now give an intuitive explanation of the Teichm\"{u}ller density. Holomorphic functions are functions which satisfy $\bar{\delta}f = 0$, so in essence $\lambda_A$ is measuring how nearly holomorphic a function can be while attaining the prescribed values at $z$ and on the boundary. The definition of the Teichm\"{u}ller density arises most naturally in the context of holomorphic motions. Given any closed set $E$ in the extended complex plane $\hat{\mathbb{C}}$, a holomorphic motion $h_t(w)$ is a function from $\mathbb{D} \times E \longrightarrow \hat{\mathbb{C}}$ which satisfies the following properties. \begin{itemize} \item[i)]{$t \longrightarrow h_t(w)$ is holomorphic on $\mathbb{D}$ for every fixed $w \in E$.} \item[ii)]{$w \longrightarrow h_t(w)$ is injective on $E$ for every fixed $t \in \mathbb{D}$.} \item[iii)]{$f_0(w) = w$ for every $z \in E$.} \end{itemize} Note the lack of any sort of continuity assumption on $h$ as a function of $w$. For this reason, it may be helpful to suppose first that $E$ is a finite set and that $h$ effects a simultaneous motion of the points in $E$. This motion has a complex time variable $t$, and the points in $E$ are not allowed to collide at any time. However, it is a remarkable fact that the holomorphicity in $t$ forces $h$ to satisfy strong continuity conditions in $w$ if $E$ is an infinite set. In fact, much more is true. The following statement, commonly referred to as the $\lambda-lemma$, was first proved by Slodkowski, although a weaker version had been proved earlier by Sullivan and Thurston (see \cite{fyun} for a complete account). \begin{lemma} Suppose $E$ is a closed set, and $h_t(w): \mathbb{D} \times E \longrightarrow \hat{\mathbb{C}}$ is holomorphic motion. Then there is a holomorphic motion $\tilde{h}_t(w): \mathbb{D} \times \hat{\mathbb{C}} \longrightarrow \hat{\mathbb{C}}$ such that $\tilde{h}$ agrees with $h$ on $\mathbb{D} \times E$. For fixed $t$, the function $w \longrightarrow \tilde{h}_t(w)$ is a quasiconformal homeomorphism from $\hat{\mathbb{C}}$ to $\hat{\mathbb{C}}$. \end{lemma} Note that nothing is said about uniqueness, and in general there will be many possible extensions of a given $h$. Suppose $E=U^c \bigcup \{z\}$, and suppose further that we have a holomorphic motion $h_t(w)$ such that $\frac{d}{dt} h_t(w)\Big|_{t=0} = 0$ for all $w \in U^c$ and $\frac{d}{dt} h_t(z)\Big|_{t=0} = 1$. Let $\tilde{h}$ be an extension of this holomorphic motion to all of $\hat{\mathbb{C}}$ guaranteed by Slodkowski's Theorem, and let us consider $\phi:=\frac{d^2}{d\bar{w}dt} \tilde{h}_t(w)\Big|_{t=0}$. Taking the $t$ derivative first and setting $V(w)=\frac{d}{dt}\tilde{h}_t(w)\Big|_{t=0}$, we obtain $\phi = \frac{dV}{d\bar{w}}$. On the other hand, $\tilde{h}_t(w)$ is quasiconformal in $w$ and thus has a Beltrami coefficient $\mu_t(z)$ such that $\frac{d\tilde{h}}{d\bar{w}} = \mu_t \frac{d\tilde{h}}{dw}$. We then have \begin{equation} \label{} \frac{d\mu_t}{dt} = \frac{\frac{d\tilde{h}}{dw}\frac{d^2\tilde{h}}{dtd\bar{w}}-\frac{d\tilde{h}}{d\bar{w}}\frac{d^2\tilde{h}}{dtdw}}{\Big(\frac{d\tilde{h}}{dw}\Big)^2} \end{equation} Note that $\frac{d\tilde{h}}{d\bar{w}}(w)\Big|_{t=0}=0$ and $\frac{d\tilde{h}}{dw}(w)\Big|_{t=0}=1$ for all $w$, so that $\phi = \frac{d^2}{d\bar{w}dt} \tilde{h}_t(w)\Big|_{t=0} = \frac{d\mu_t}{dt}\Big|_{t=0}$. Comparing our two expressions for $\phi$, we see $\frac{dV}{d\bar{w}}(z) = \frac{d\mu_t}{dt}(z)\Big|_{t=0}$. Thus, the Teichm\"{u}ller density measures the minimal rate of change of the Beltrami coefficient of all quasiconformal homeomorphisms associated to a holomorphic motion which fixes the boundary of $U$ and moves $z$ with unit velocity at time $0$. It stands to reason that points closer to the boundary, in whatever sense, require a more violent holomorphic motion in order to move (while the boundary remains fixed) than those farther away. This results in a larger value of $\lambda_U$, as is the case with the hyperbolic metric $\rho_U$. \hspace{.25in} This density is well suited to our problem concerning the rate of convergence of densities on domains, as we can use known moduli of continuity on vector fields associated to holomorphic motions to our advantage. Suppose that a sequence of domains $U_n$ converges in boundary to a domain $U$. The following theorem shows that the Teichm\"{u}ller metric on $U_n$ converges to the Teichm\"{u}ller metric on $U$ with speed $\log(\log(\frac{1}{H(\delta U_n,\delta U)}))$. \begin{theorem} \label{big} Suppose that $U$ is a bounded domain, and that $K$ is a compact subset of $U$. Then there is a constant $C$ such that if $W$ is a domain with $H(\delta U,\delta W) \leq \varepsilon$ then $|\lambda_U (w) - \lambda_W (w)| \leq C (\log(\log(1/\varepsilon)))^{-1}$ for all $w \in K$, provided that $\varepsilon$ is sufficiently small. The constant $C$ depends on $d(K,\delta U)$ and the diameter of $U$. \end{theorem} {\bf Proof:} Let $\varepsilon < \varepsilon_o<e^{-1}$ where $\varepsilon_o$ is sufficiently small so that the condition $d(\delta U,\delta W) \leq \varepsilon_o$ forces $W$ to contain $K$. Let $d_U (z) = d(z,\delta U)$. Given $w \in K$, let $V$ denote a differentiable vector field on $U$ with $V(w) = 1$, $V(z) = 0$ for $z \in U^c$, and $||\bar{\delta}V||_{\infty} = \lambda_U (w)$. We will use $V$ to construct a vector field on $W$ which will give us a bound for $\lambda_W (w)$. The problem, of course, is that $V(z)$ is not necessarily 0 on $W^c$, so we must alter it at the boundary in a way that doesn't affect the norm of the $\bar{\delta}$ derivative very much. We can assume, after multiplying the entire picture by a constant if necessary, that $d(\delta U,K) \geq 1/2$. Let \begin{equation} j(x)= \left \{ \begin{array}{ll} 0, & \qquad \mbox{if { }} 0 \leq x \leq \varepsilon, \\ (\log(\log(1/\varepsilon)))^{-1}\int_{\varepsilon}^x \frac{1}{t\log(1/t)} dt, & \qquad \mbox{if { }} \varepsilon < x \leq e^{-1},\\ 1, & \qquad \mbox{if { }} e^{-1} < x. \end{array} \right . \end{equation} Then \begin{equation} j'(x) = (x\log(1/x))^{-1}(\log(\log(1/\varepsilon)))^{-1}1_{[\varepsilon,e^{-1}]}(x)\end{equation} in the distributional sense. Let $\chi (z) = j(d_U(z))$. Since $d_U(z)$ is Lipshitz, it has distributional derivatives of norm at most 1, and thus \begin{equation} \label{nice}|\bar{\delta}\chi(z)|\leq \Big(d_U (z)\log(1/d_U (z))\log(\log(1/\varepsilon))\Big)^{-1} . \end{equation} Note also that $0 \leq \chi \leq 1$, with $\chi(w) = 1$ and $\chi(z) = 0$ whenever $d_U(z)< \varepsilon$. Let $\hat{V}(z) = V(z)\chi(z)$. Since $H(\delta U,\delta W) \leq \varepsilon$ we see that $\hat{V}(z)$ is 0 on $W^c$ and 1 at $w$, so it gives an upper bound for $\lambda_W(w)$. We obtain \begin{equation} \bar{\delta}\hat{V} = (\bar{\delta}V)\chi + V(\bar{\delta}\chi) \end{equation} so that \begin{equation} \label{cool} |\bar{\delta}\hat{V}| \leq |(\bar{\delta}V)| + |V(\bar{\delta}\chi)| . \end{equation} We have the estimate $|V(z)| \leq |C d_U(z) \log(1/d_U(z))|$ from Theorem 7 of Chapter 3 in \cite{fnbook}, where $C$ here depends on $d(K,\delta U)$ and the diameter of $U$. In light of this and (\ref{nice}), we see that (\ref{cool}) is bounded above by \begin{equation} ||V||_\infty + C(\log(\log(1/\varepsilon)))^{-1} . \end{equation} Thus, $\lambda_W (w) \leq \lambda_U (w) + C(\log(\log(1/\varepsilon)))^{-1}$ for all $w \in K$. Interchanging the roles of $U$ and $W$ in the above argument gives the reverse inequality, and shows that $|\lambda_U (w) - \lambda_W (w)| \leq C (\log(\log(1/\varepsilon)))^{-1}$. {\hfill $\Box$ \bigskip} \section{Rates of convergence for the hyperbolic density.} As mentioned in Section \ref{tm}, the Teichm\"{u}ller and hyperbolic densities coincide, up to a constant, on simply connected domains. The following is therefore a corollary to Theorem \ref{big}. \begin{cor} Suppose that $U$ is a simply connected, bounded domain, and that $K$ is a compact subset of $U$. Then there is a constant $C$ such that if $W$ is a simply connected domain with $H(\delta U,\delta W) \leq \varepsilon$ then $|\rho_U (z) - \rho_W (z)| \leq C (\log(\log(1/\varepsilon)))^{-1}$ for all $z \in K$, provided that $\varepsilon$ is sufficiently small. The constant $C$ depends on $d(K,\delta U)$ and the diameter of $U$. \end{cor} {\bf Remark:} The boundedness requirement may be relaxed in certain cases. For instance, if $U$ is unbounded but $U^c$ contains an open set containing a point $p$, we may apply a M\"{o}bius inversion mapping $p$ to $\infty$. The image under this map is a bounded domain, and the spherical metric is preserved by the inversion up to a constant and is equivalent to the Euclidean metric in the bounded image. Thus, the corollary may be applied, with $H(\delta U,\delta W)$ now being measured in the spherical metric. \vspace{12pt} It was also mentioned in Section \ref{tm} that the Teichm\"{u}ller and hyperbolic densities coincide on the largest possible hyperbolic domain, the thrice punctured plane $\hat{\mathbb{C}} \backslash \{a,b,c\}$. Thus, we may obtain a similar result if $U_n$ and $U$ are thrice punctured planes. However, as the hyperbolic density of a thrice punctured plane is explicitly computable we may obtain a far better rate of convergence when $U_n \longrightarrow U$ in boundary, namely $H \log(\frac{1}{H})$, where $H=H(\delta U_n,\delta U)$. Let the hyperbolic metric on $\hat{\mathbb{C}}\backslash \{a,b,c\}$ be denoted $\rho_{a,b,c}(z)$. \begin{theorem} \label{jets} Let $U=\mathbb{C}\backslash \{a,b,c\}$ and let $K$ be a compact set in $U$. Then there is a constant $C$ such that $|\rho_{a,b,c}(z) - \rho_{a',b',c'}(z)|<C\varepsilon \log(\frac{1}{\varepsilon})$ for all $z \in K$ whenever $H(\{a,b,c\}, \{a',b',c'\}) < \varepsilon$ and $\varepsilon$ is sufficiently small. $C$ depends on $\sup_{w \in K} \{|w-a|,|w-b|,|w-c|\}, d(K,\{a,b,c\}),|a-b|, |b-c|,$ and $|a-c|$. \end{theorem} {\bf Proof:} With no loss of generality we may assume that $|a-a'|,|b-b'|, |c-c'| \leq \varepsilon.$ The triangle inequality \begin{eqnarray} \label{} && \nonumber |\rho_{a,b,c}(z)-\rho_{a',b',c'}(z)| \leq |\rho_{a,b,c}(z)-\rho_{a,b,c'}(z)| \\ \nonumber && \hspace{1cm} +|\rho_{a,b,c'}(z)-\rho_{a,b',c'}(z)|+|\rho_{a,b',c'}(z)-\rho_{a',b',c'}(z)| \end{eqnarray} implies that we may assume that $a=a'$ and $b=b'.$ \cite{fnbook} contains a proof of the following formula: \begin{equation} \label{ref} \frac{1}{\rho_{a,b,c}(z)} = \frac{1}{\pi} \int \! \! \int_{\mathbb{C}} \frac{|(z-a)(z-b)(z-c)|}{|(w-a)(w-b)(w-c)(w-z)|} dA(w). \end{equation} Using this formula together with the corresponding expression for $\rho_{a,b,c'}(z)$ we have \begin{eqnarray} \label{wes} && |\rho_{a,b,c}(z) - \rho_{a,b,c'}(z)| = \frac{1}{\pi}\rho_{a,b,c}(z) \rho_{a,b,c'}(z) \\ \nonumber && \hspace{1cm} \times \int \! \! \int_{\mathbb{C}} \frac{|(z-a)(z-b)|}{|(w-a)(w-b)(w-z)|} \Big( \frac{|z-c|}{|w-c|} -\frac{|z-c'|}{|w-c'|} \Big)dA(w). \end{eqnarray} Let $M = \sup_{w \in K} \{|w-a|,|w-b|,|w-c|\}$, $d = d(K,\{a,b,c\})$, and $m=\min\{|a-b|, |b-c|,|a-c|, d\}$. Now \begin{equation} \label{wel} \frac{|z-c|}{|w-c|} -\frac{|z-c'|}{|w-c'|}\leq \frac{|z-c|}{|w-c|} -\frac{|z-c'|}{|w-c|}+\frac{|z-c'|}{|w-c|} -\frac{|z-c'|}{|w-c'|}. \end{equation} Thus, \begin{equation} \label{dir} |\rho_{a,b,c}(z) - \rho_{a,b,c'}(z)| \leq \rho_{a,b,c}(z) \rho_{a,b,c'}(z) (I + II) \end{equation} where \begin{equation} I=\frac{1}{\pi}\int \! \! \int_{\mathbb{C}} \frac{|(z-a)(z-b)(c-c')|}{|(w-a)(w-b)(w-c)(w-z)|}dA(w)\end{equation} and \begin{equation} II=\frac{1}{\pi}\int \! \! \int_{\mathbb{C}} \frac{|(z-a)(z-b)(z-c')(c-c')|}{|(w-a)(w-b)(w-c)(w-c')(w-z)|}dA(w) .\end{equation} Note that \begin{equation} \rho_{a,b,c}(z) \leq \rho_{D(z,d)}(z)=\frac{1}{d}\end{equation} and similarly \begin{equation}\rho_{a,b,c'}(z) \leq \frac{1}{d-\varepsilon}.\end{equation} Furthermore, \begin{equation} I=\frac{1}{\pi}\int \! \! \int_{\mathbb{C}} \frac{|(z-a)(z-b)(z-c)|}{|(w-a)(w-b)(w-c)(w-z)|}\frac{|c-c'|}{|z-c|}dA(w) \leq \frac{\varepsilon}{d}\frac{1}{\rho_{a,b,c}(z)}.\end{equation} We will estimate II by considering two different regions. Let $A=\{w\in \mathbb{C}: |w-c'|>\frac{d}{2}\}$ and let $B$ be the complement of $A.$ Then \begin{equation} II=\frac{1}{\pi}\int \! \! \int_{A} \frac{|(z-a)(z-b)(z-c')(c-c')|}{|(w-a)(w-b)(w-c)(w-c')(w-z)|}dA(w)+\end{equation} $$\frac{1}{\pi}\int \! \! \int_{B} \frac{|(z-a)(z-b)(z-c')(c-c')|}{|(w-a)(w-b)(w-c)(w-c')(w-z)|}dA(w).$$ We have \begin{equation} \frac{1}{\pi}\int \! \! \int_{A} \frac{|(z-a)(z-b)(z-c')(c-c')|}{|(w-a)(w-b)(w-c)(w-c')(w-z)|}dA(w)=\end{equation} $$\frac{1}{\pi}|c-c'|\int \! \! \int_{A} \frac{|(z-a)(z-b)(z-c)|}{|(w-a)(w-b)(w-c)(w-z)|}\frac{|z-c'|}{|w-c'||z-c|}dA(w)$$ $$\leq \frac{2\varepsilon (M+\varepsilon)}{d^2\rho_{a,b,c}(z)},$$ and for the integral over $B$ we obtain \begin{equation} \frac{1}{\pi}\int \! \! \int_{B} \frac{|(z-a)(z-b)(z-c')(c-c')|}{|(w-a)(w-b)(w-c)(w-c')(w-z)|}dA(w)=\end{equation} $$\frac{1}{\pi}\int \! \! \int_{B} \frac{|(c'-a)(c'-b)(c'-c)|}{|(w-a)(w-b)(w-c)(w-c')|}\frac{|(z-a)(z-b)(z-c')|}{|(c'-a)(c'-b)(w-z)|}dA(w)$$ $$\leq \frac{M^2 (M+\varepsilon)}{(\frac{d}{2}-\varepsilon)(m-\varepsilon)^2\rho_{a,b,c}(c')}$$ where we have used \begin{equation} \label{ref1} \frac{1}{\rho_{a,b,c}(c')} = \frac{1}{\pi} \int \! \! \int_{\mathbb{C}} \frac{|(c'-a)(c'-b)(c'-c)|}{|(w-a)(w-b)(w-c)(w-c')|} dA(w) . \end{equation} Now let $f$ be the Mobius transformation \begin{equation} f(z)=\frac{(z-c)(a-b)}{(z-b)(a-c)}.\end{equation} We have $f(c)=0, f(b)=\infty$ and $f(a)=1.$ Thus \begin{equation} \rho_{a,b,c}(z)=\rho_{0,1}(f(z))|f'(z)|=\rho_{0,1}(f(z))\frac{|(a-b)(b-c)|}{|(a-c)(z-b)^2|}\end{equation} where we are using the notation $\rho_{0,1}(z)$ as a shorthand for $\rho_{0,1,\infty}(z)$. For $z \in W,$ \begin{equation} \frac{|(a-b)(b-c)|}{|(a-c)(z-b)^2|}\geq \frac{m^2}{2Md^2}. \end{equation} Note that $|f(z)| = \frac{|(z-a)(a-b)|}{|(z-b)(a-c)|} \leq \frac{2M^2}{dm}$. Theorem 14.3.1 in \cite{nlbook} shows then that \begin{equation} \label{} \rho_{0,1}(f(z)) \geq \frac{1}{2|\frac{2M^2}{dm}|\log|\frac{2M^2}{dm}| + 10 \frac{2M^2}{dm}}. \end{equation} Thus, $\rho_{a,b,c}(z)$ is bounded below by a constant depending on $m,M,d$, and we conclude that \begin{equation} \frac{1}{\pi}\int \! \! \int_{A} \frac{|(z-a)(z-b)(z-c')(c-c')|}{|(w-a)(w-b)(w-c)(w-c')(w-z)|}dA(w)\leq C\varepsilon.\end{equation} Similarly, \begin{equation} \rho_{a,b,c}(c')=\rho_{0,1}(f(c'))|f'(c')|=\rho_{0,1}(f(z))\frac{|(a-b)(b-c)|}{|(a-c)(c'-b)^2|}.\end{equation} We have \begin{equation} \frac{|(a-b)(b-c)|}{|(a-c)(c'-b^2)|}\geq \frac{m^2}{2M(2M+\varepsilon)^2}\end{equation} and \begin{equation} f(c')=\frac{|(a-b)(c'-c)|}{|(a-c)(c'-b)|}\leq \frac{2M\varepsilon}{m(m-\varepsilon)}<\frac{1}{2}\end{equation} for sufficiently small $\varepsilon.$ Corollary 14.4.1 in \cite{nlbook} implies \begin{equation} \frac{1}{\rho_{01}(f(c'))} \leq 17 |f(c')|\log (\frac{1}{|f(c')|}) \leq C \varepsilon \log (\frac{1}{\varepsilon}).\end{equation} Thus, we have \begin{equation} \frac{1}{\pi}\int \! \! \int_{B} \frac{|(z-a)(z-b)(z-c')(c-c')|}{|(w-a)(w-b)(w-c)(w-c')(w-z)|}dA(w)\leq C\varepsilon\log (\frac{1}{\varepsilon}),\end{equation} and this proves the theorem. {\hfill $\Box$ \bigskip} In the prior theorem we have not allowed any of $a,b,c$ to be $\infty$. If we allow one of the points to be $\infty$, using the notation $\rho_{a,b}$ in place of $\rho_{a,b,\infty}$, we can obtain the following. \begin{theorem} \label{suck} Let $U = \mathbb{C} \backslash \{a,b\}$ and let $K$ be a compact set in $U$. Then there is a constant $C$ such that $|\rho_{a,b}(z)-\rho_{a',b'}(z)| < C \varepsilon \log(\frac{1}{\varepsilon})$ for all $z \in K$ whenever $H(\{a,b\},\{a',b'\})<\epsilon$. The constant $C$ depends on $\sup_{w \in K} \{|w-a|,|w-b|\}$, $d(K,\{a,b\})$ and $|a-b|$. \end{theorem} {\bf Proof:} Let $\gamma$ be a line segment joining $K$ and $\{a,b\}$ with the length of $\gamma$ equal to $d(K,\{a,b\})$. Let $r$ be the midpoint of $\gamma$. For any point $z$ in $K$ the M\"{o}bius transformation $g(z)=\frac{1}{(z-r)}$ satisfies \begin{equation} \label{} \rho_{a,b}(z)=\rho_{1/(a-r),1/(b-r),0}(1/(z-r))\frac{1}{|z-r|^2} \end{equation} and \begin{equation} \label{} \rho_{a',b'}(z)=\rho_{1/(a'-r),1/(b'-r),0}(1/(z-r))\frac{1}{|z-r|^2}. \end{equation} Theorem \ref{suck} now follows from Theorem \ref{jets} applied to the compact set $\{\frac{1}{(z-r)}: z \in K\}$, using the the point $\frac{1}{(z-r)}$ in place of $z$. {\hfill $\Box$ \bigskip} In the case of the the unit disc, we can obtain a $H(\delta U_n,\delta \mathbb{D})$ rate of convergence. \begin{theorem} If $K$ is a compact subset of $\mathbb{D}$, then there is a constant $C$ depending on $K$ such that $|\rho_{\mathbb{D}}(z)-\rho_W(z)| \leq C\varepsilon$ for all $z$ in $K$ whenever $W$ contains $K$ and $H(\delta \mathbb{D}, \delta W) \leq \varepsilon$ for $\varepsilon$ sufficiently small. \end{theorem} {\bf Proof:} For $r>0$ let $r\mathbb{D} = \{|z|<r\}$. The hyperbolic density on $r\mathbb{D}$ is well known to be given by $\rho_{r\mathbb{D}}(z) = \frac{r}{r^2-|z|^2}$ (see \cite{fnbook}). Thus, \begin{equation} \label{} \begin{split} |\rho_{r\mathbb{D}}(z)-\rho_{\mathbb{D}}(z)| & = \frac{|r(1-|z|^2) - (r^2-|z|^2)|}{(1-|z|^2)(r^2-|z|^2)} \\ & \leq \frac{(r|1-r|+ |z|^2|1-r|)}{(\min (1,r^2) - |z|^2)^2}\leq C|1-r| \end{split} \end{equation} for $|z|$ uniformly bounded below $r$. The result now follows by the monotonicity property of the hyperbolic density, as when $H(\delta \mathbb{D}, \delta W) \leq \varepsilon$ we must have $(1-\varepsilon)\mathbb{D} \subseteq W \subseteq (1+\varepsilon)\mathbb{D}$. {\hfill $\Box$ \bigskip} {\bf Remark:} The examples $\{|z|<1+\varepsilon\}$ and $\{|z|<1-\varepsilon\}$ show that this rate of convergence can not be improved. \vspace{12pt} Though not directly related to the results given in this section, we would be remiss if we did not mention the one prior result we have seen concerning the rate of convergence of the hyperbolic density. In \cite{beard} the rate of convergence of $\rho_U(x)$ was determined for $x \in R$ and $U$ of the form $\{|x| < l, |y| < \frac{\pi}{2}\}$ for changing values of $l$. \section{Remarks on the three-point density.} Suppose $U$ is an arbitrary domain in $\mathbb{C}$. If $a,b,c \in U^{c}$, then $\rho_{a,b,c}(z) \leq \rho_U(z)$ for all $z \in U$ by the monotonicity of the hyperbolic density. We can define a new density on $U$ by setting \begin{equation} h_U(z) = \sup_{a,b,c \in U^{c}} \rho_{a,b,c}(z) . \label{8} \end{equation} This was first done in \cite{fnart}, and we shall refer to this quantity as the {\it three-point density}. We clearly have $h_U \leq \rho_U$, and it is true, though less obvious, that equality holds only when $U$ is itself the thrice punctured sphere. It was shown initially in \cite{fnart} that there is a positive universal constant $C$ such that \begin{equation} \label{dars} \rho_U \leq Ch_U, \end{equation} so that the two densities are equivalent. In \cite{minma}, it was shown that $h_U$ is a continuous density which is M\"{o}bius invariant. \cite{suvu} gave an explicit constant for \rrr{dars}, and \cite{bets} worked to improve the constant and also calculated $h_\mathbb{D}$. In relation to the convergence of densities, we have the following proposition. \begin{prop} If $z \in U$ and $U_n \longrightarrow U$ in boundary, then $h_{U_n}(z) \longrightarrow h_U(z)$. \end{prop} Proof: If $a,b,c \in U^c$ we can choose $a_n,b_n,c_n \in U_n^c$ which converge to $a,b,c$ respectively. By Lemma \ref{old}, \begin{equation} \rho_{a_n,b_n,c_n}(z) \longrightarrow \rho_{a,b,c}(z) . \label{9} \end{equation} It follows from this that \begin{equation} \underline{\lim}h_{U_n}(z) \geq h_U(z) . \label{10} \end{equation} For the reverse inequality, we can choose $a_n,b_n,c_n \in U_n^c$ such that \begin{equation} \overline{\lim} \rho_{a_n,b_n,c_n}(z) = \overline{\lim} h_{U_n}(z) . \label{10a} \end{equation} After passing to subsequences several times if necessary we may assume $a_n,b_n,c_n \longrightarrow a,b,c$ respectively. Since $U_n \longrightarrow U$ we can choose $a_n',b_n',c_n' \in U^c$ close to $a_n,b_n,c_n$, such that $a_n',b_n',c_n' \longrightarrow a,b,c$. $U^c$ is closed, so $a,b,c \in U^c$. If $a,b,c$ are distinct, then applying Lemma \ref{old} we see that \begin{equation} \overline{\lim}h_{U_n}(z) \leq h_U \label{11}(z) \end{equation} completing the proof. It remains only to see that $a,b,c$ must be all distinct, since if not then we would approach a pole of order two in \rrr{ref} as $n \longrightarrow \infty$. This would force $\rho_{a_n,b_n,c_n}(z) \longrightarrow 0$, contradicting (\ref{10}) and \rrr{10a}. {\hfill $\Box$ \bigskip} It would seem that Theorem \ref{jets} was ideally suited for deducing a rate of convergence result for $h_{U_n}$. In fact, it was shown in \cite{minma} that the supremum in \rrr{8} is always attained for some triple $a,b,c$ in $\delta U$, so if $U$ is bounded we may assume $a,b,c$ are bounded in \rrr{8} as well. The difficulty, however, lies in the fact that the constant in Theorem \ref{jets} depends in part on $(\min\{|a-b|, |b-c|,|a-c|\})^{-1}$. If a point is fixed in $U$ then there are $a,b,c \in \delta U$ such that $h_U(z) = \sup_{a,b,c \in U^{c}} \rho_{a,b,c}(z)$. We may then apply Theorem \ref{jets} to obtain \begin{equation} \label{} h_{U_n}(z) \geq h_U(z) - C \sqrt{H(\delta U, \delta U^n)} \end{equation} for $H(\delta U, \delta U^n)$ sufficiently small. However, different points in a compact set $K$ determine different optimal triples of points $a,b,c$, and we do not currently have a way to bound $\min\{|a-b|, |b-c|,|a-c|\}$ from below. For the argument to show $\limsup h_{U_n}(z) \leq h_U(z)$, we know that $a_n,b_n,c_n$ converge, but we do not know how far the limit points are from each other. Obtaining a theorem on the rate of convergence of $h_U$ would seem therefore to necessitate understanding how the optimal points $a,b,c$ are situated in the plane for given $z$ and $U$. \section{Further questions} It may be of interest for applications to explicitly calculate the constants in the results given above. It would also be interesting to know whether the rates of convergence given are the best possible. Except where stated, we do not know whether this is the case. Perhaps there are results similar to the ones in this paper for any of a number of other densities, for instance the Carath\'{e}odory density or Kobayashi density in higher dimensions. Of course, finding an analog of Corollary 1 for domains which are not simply connected would be desirable as well. \section{Acknowledgements} We would like to thank Fred Gardiner for many helpful conversations, as well as an anonymous referee for a careful reading. \def\noopsort#1{} \def\printfirst#1#2{#1} \def\singleletter#1{#1} \def\switchargs#1#2{#2#1} \def\bibsameauth{\leavevmode\vrule height .1ex depth 0pt width 2.3em\relax\,} \makeatletter \renewcommand{\@biblabel}[1]{\hfill#1.}\makeatother \bibliographystyle{alpha}
\section{Introduction} Lorentz violating models are being object of intensive investigation since the beginning of the nineties \cite{kostelecky1}-\cite{ferreira}. The Standard Model Extension (SME) \cite{kostelecky1}-\cite{coleman2} proposes a wide range of possibilities in this context, concerning both the gauge and fermion sectors. There are two particularly interesting terms in the photon sector of the SME: the CPT-odd term of Carroll-Field-Jackiw \cite{jackiw} and a CPT-even term which is quadratic in the field strength \cite{kostelecky2}. In the context of the SME, there have been great interest in the quantum induction of the Carroll-Field-Jackiw term \cite{CS1}-\cite{scarp1}. On the other hand, the radiative generation of the CPT-even one (an aether-like term \cite{carroll}) has been investigated in the context of effective models which include a Lorentz violating nonminimal (magnetic) coupling \cite{mgomes1}, \cite{mgomes2}, \cite{scarp-NM}. The perturbative generation of higher derivative Lorentz-breaking terms from the nonminimal coupling has also been investigated \cite{petrov}. This kind of model has been considered before in papers \cite{Belich0}-\cite{Belich3} in the context of Relativistic Quantum Mechanics. It was used in the calculation of corrections to the Hydrogen spectrum, from which very stringent bounds have been set up in the magnitude of the Lorentz violating parameter \cite{Belich2}. It was also used to study the magnetic moment generation from the nonminimal coupling, since a tiny magnetic dipole moment of elementary neutral particles might signal Lorentz symmetry violation \cite{Belich3}. In \cite{mgomes1} and \cite{mgomes2} the magnetic coupling has been considered in the place of the minimal coupling in a gauge violating model. The coefficient of the induced aether-like term was shown to be regularization dependent. The inclusion of the minimal coupling together with the magnetic one was considered in paper \cite{scarp-NM}. In this case, if a gauge invariant regularization prescription is used, there is no quantum induction of Lorentz-violating terms in the gauge sector (CPT-even or -odd) at one-loop order. However, a gauge violating regularization technique followed by a symmetry restoring counterterm in the Lorentz symmetric sector opens the possibility for such inductions. In this paper, another kind of nonminimal coupling is considered, in which the background vector $b_\mu$ appears coupled to the gauge field by means of a term of chiral character. It is shown that, in this case, there is no induction of CPT-odd terms. On the other hand, the one-loop photon self-energy will include an aether-like CPT-even part, with a divergent coefficient. This shows that this aether-like part must be included from the beginning in the gauge sector. This fact can be understood in terms of the nonrenormalizability of the model. If higher order contributions are considered, this term is just the first one of an infinite list that should be included from the beginning. We discuss this fact is the context of an effective model, in which the determination of a physical cutoff is an essential step. This paper is organized as follows: the second section is dedicated to a general discussion on nonminimal coupling models; the third section presents the one-loop calculation of the photon self-energy for the model with a chiral nonminimal coupling; the conclusions are drawn in section four. \section{Discussion on QED with nonminimal couplings} The action of a QED model with a nonminimal coupling is written as \be \Sigma = \int d^4x \left\{ -\frac 14 F_{\mu \nu} F^{\mu \nu} + \bar \psi \left( i \parbruto - m - e \Abruto - g \varepsilon^{\mu \nu \alpha \beta} \Gamma_\beta b_\mu F_{\nu \alpha} \right) \psi \right\}, \label{action} \ee where $\Gamma_\beta=\gamma_\beta$ for the simple magnetic coupling and $\Gamma_\beta=\gamma_5 \gamma_\beta$ for the chiral nonminimal coupling. Conventional QED is recovered in the limit $g \to 0$. The action of equation (\ref{action}) describes a gauge invariant model with some interesting classical features as investigated in \cite{Belich0}. Indeed, it is shown for the case with $\Gamma_\beta=\gamma_\beta$, that the 3-vector $\vec b$ plays the role of a kind of magnetic dipole moment ($\vec \mu= g \vec b$). Besides, in this case there is an induction of a Aharonov-Casher (A-C) effect. On the other hand, the chiral nonminimal coupling contributes to the interaction energy without inducting a A-C phase. We present now a general discussion on the one-loop correction to the photon self-energy for a QED with nonminimal coupling. First, a great simplification in the calculations is obtained if we write $B^\beta=\varepsilon^{\mu \nu \alpha \beta} b_\mu F_{\nu \alpha}$. In the case of the simple nonminimal coupling, if the purpose is only the one-loop calculation of the vacuum polarization tensor, it is yet possible to define the field $\tilde A_\mu=e A_\mu+g B_\mu$, so that the new lagrangian density for the fermion sector can be written as \be {\cal L}_{\psi}= \bar \psi \left( i \parbruto - m - \tilde \Abruto \right) \psi. \ee In terms of $\tilde A_\mu$, the one-loop correction to the photon two-point function is identical to the one for the conventional QED. In momentum space, we have \be T^{\mu\nu}(p)=\text{tr}\int_k{\gamma^\nu s(p+k)\gamma^\mu s(k)}, \label{T} \ee in which $s(k)$ is the fermion propagator and $\int_k$ stands for $\int d^4k/\left(2 \pi\right)^4$. The corrections to the photon sector in the lagrangian density will be given by \bq &&-\frac 12 \tilde A^\mu T_{\mu \nu}(x) \tilde A^\nu \nonumber \\ &&=-\frac 12 A^\mu e^2 T_{\mu \nu}(x) A^\nu -\frac 12 A^\mu 2 eg T_{\mu \nu}(x) B^\nu -\frac 12 B^\mu g^2 T_{\mu \nu}(x) B^\nu. \label{magcoupling} \eq In this case in which chirality is absent, treated in \cite{scarp-NM}, the discussion of the induced Lorentz-violating terms can be performed in general grounds. A general expression for $T_{\mu \nu}$ obtained by means of some regularization technique, not necessarily gauge invariant, compatible with its Lorentz structure is \be T_{\mu \nu}=\left(p_\mu p_\nu -p^2 g_{\mu \nu} \right) \Pi (p^2) + \alpha m^2 g_{\mu \nu} + \beta p_\mu p_\nu, \label{Tmunu} \ee in which $\alpha$ and $\beta$ are dimensionless constants. In eq. (\ref{magcoupling}), the first term is the traditional QED one. The second is a Lorentz-violating CPT-odd Chern-Simons-like term, which in the on-shell limit is given by \be {\cal L}_{CS}=-\alpha eg m^2 \varepsilon^{\mu \nu \alpha \beta} b_\mu A_\nu F_{\alpha \beta}. \ee The third term in eq. (\ref{magcoupling}) is a CPT-even term, which in the on-shell limit is written as \be {\cal L}_{even}=-\alpha m^2 g^2 b^2 F_{\mu \nu}F^{\mu \nu}+2 \alpha m^2 g^2 \left(b_\mu F^{\mu \nu} \right)^2, \label{Even} \ee where \be B^\mu B_\mu= 2 b^2 F_{\mu \nu}F^{\mu \nu}-4 \left(b_\mu F^{\mu \nu} \right)^2 \ee has been obtained with the help of some properties of the L\'evi-Civit\`a tensor. We recognize in the second term of ${\cal L}_{even}$ the Lorentz-violating aether term. A comment is in order. The value of the constant $\alpha$ is determined by the regularization procedure used in the calculation. If a regularization technique which preserves gauge invariance of the original QED is used, $\alpha$ will be null. However, it is always possible to choose a gauge breaking procedure and then restore the symmetry by means of a non-symmetric counterterm. In such case, since the Lorentz violating and Lorentz preserving parts are independent on each other, the Lorentz breaking terms would survive. Such procedure is equivalent to use different regularization techniques in different sectors. Nevertheless, the natural framework is using an unique regularization in the calculation of integrals which contribute to the same amplitude. In this case, the calculation with a gauge preserving method would not induce, at one-loop order, these two Lorentz-violating terms. The discussion above will not apply to the case of chiral nonminimal interaction, as it will be presented in the next section. \section{Quantum corrections to the QED with chiral nonminimal coupling} We now turn our attention to the the one-loop quantum corrections to the photon sector in a model with a nonminimal interaction with a chiral character. We now have the following lagrangian density for the fermion sector: \be {\cal L}_{\psi}= \bar \psi \left( i \parbruto - m - e \Abruto -g \gamma_5 \Bbruto \right) \psi. \ee This fermion lagrangian in terms of the vector and axial-vector fields $A_\mu$ and $B_\mu$, in the case in which the fields do not depend on each other, has been vastly investigated (see, for example, \cite{Bardeen} and \cite{Bonora}). The one-loop corrections to the photon sector will be given by \be -\frac 12 A^\mu e^2 T^{VV}_{\mu \nu}(x) A^\nu -\frac 12 A^\mu 2 eg T^{AV}_{\mu \nu}(x) B^\nu -\frac 12 B^\mu g^2 T^{AA}_{\mu \nu}(x) B^\nu, \label{chicoupling} \ee where the upper indices $A$ and $V$ refer to the axial and vectorial vertices, so that $T^{VV}_{\mu \nu}=T_{\mu \nu}$. In momentum space, we have \be T^{AV}_{\mu \nu}(p)=\int_k{\text{tr}\left\{\gamma^\nu s(p+k)\gamma_5\gamma^\mu s(k)\right\}} \label{TAV} \ee and \be T^{AA}_{\mu \nu}(p)=\int_k{\text{tr}\left\{\gamma_5\gamma^\nu s(p+k)\gamma_5\gamma^\mu s(k)\right\}}. \ee The second term in (\ref{chicoupling}), which would be CPT-odd, is actually identically null, since in eq. (\ref{TAV}), we have in the numerator \bq &&\text{tr}\left\{\gamma_5\gamma^\rho (\pbruto + \kbruto +m)\gamma^\mu (\kbruto+m)\right\}= \text{tr}\left\{\gamma_5\gamma^\rho (\pbruto + \kbruto)\gamma^\mu \kbruto\right\} + \nonumber \\ &&+m^2 \text{tr}\left\{\gamma_5\gamma^\rho \gamma^\mu \right\} + m\text{tr}\left\{\gamma_5\gamma^\rho (\pbruto + \kbruto)\gamma^\mu \right\} +m\text{tr}\left\{\gamma_5\gamma^\rho \gamma^\mu \kbruto\right\} \nonumber \\ &&=4i \varepsilon^{\rho \kappa \mu \lambda}(p+k)_\kappa k_\lambda = 4i \varepsilon^{\rho \kappa \mu \lambda}p_\kappa k_\lambda. \eq Under integration, this will vanish due to the antisymmetry of the L\'evi-Civit\`a tensor. We are left with the CPT-even term of (\ref{chicoupling}). First, we can write \be \text{tr}\left\{\gamma_5\gamma^\delta (\pbruto+\kbruto+m)\gamma_5\gamma^\rho (\kbruto+m)\right\} = \text{tr}\left\{\gamma^\delta (\pbruto+\kbruto+m)\gamma^\rho (\kbruto+m)\right\} -8m^2 g^{\rho \delta}, \ee so that \be T^{AA}_{\mu \nu}=T_{\mu \nu}-8m^2 g_{\mu \nu} I \label{aa} \ee where \be I= \int^\Lambda \frac{d^4k}{(2 \pi)^4} \frac {1}{(k^2-m^2)\left[(p-k)^2-m^2\right]} \label{I} \ee is a divergent integral and $\Lambda$ is to indicate that some regularization prescription is applied. We have to note that the contribution of the first term in (\ref{aa}) is identical to the one calculated with the nonminimal coupling without chirality. The divergent integral $I$ can be evaluated by any regularization method. In order to make the regularization independence manifest, we may write equation (\ref{I}) in a way that divergences are expressed in terms of the loop momentum only, as in Implicit Regularization \cite{papersIR}: \be I= I_{log}\left(\lambda^2\right)-b Z_0(p^2,m^2,\lambda^2), \ee where \be I_{log}\left(\lambda^2\right)=\int^\Lambda \frac{d^4k}{(2 \pi)^4} \frac {1}{(k^2-\lambda^2)^2}, \ee \be Z_0(p^2,m^2,\lambda^2)=\int_0^1 dx\,\ln{\left(\frac{p^2x(1-x)-m^2}{\left(-\lambda^2\right)}\right)}, \ee $b=i/(4 \pi)^2$ and $\lambda^2$ is an arbitrary ultraviolet mass scale. Since we are interested in the on-shell limit, we will be left with \be T^{AA}_{\mu \nu}=m^2 g_{\mu \nu} \left[-8 I_{log}\left(\lambda^2\right) +\alpha +8b \ln{\left(\frac{m^2}{\lambda^2}\right)}\right] \equiv F\left(m^2, \lambda^2\right)g_{\mu \nu}, \ee where $\alpha$ is defined in equation (\ref{Tmunu}). So, the CPT-even term will be given by \bq &&{\cal L}_{even}=-\frac 12 F\left(m^2, \lambda^2\right) B_\mu B^\mu \nonumber \\ &&= -F\left(m^2, \lambda^2\right)\left[b^2 F_{\mu \nu}F^{\mu \nu}-2 \left(b_\mu F^{\mu \nu} \right)^2 \right]. \label{result} \eq The Lorentz-violating second term above can be mapped in the CPT-even term proposed in \cite{kostelecky2}, \be {\cal L}_{even}=-\frac 14 \kappa_{\mu \nu \alpha \beta} F^{\mu \nu} F^{\alpha \beta}, \ee as long as we establish the relation \be \kappa_{\mu \nu \alpha \beta}= -2 F\left(m^2, \lambda^2\right) \left( g_{\mu \alpha} b_\nu b_\beta - g_{\nu \alpha} b_\mu b_\beta + g_{\nu \beta} b_\mu b_\alpha - g_{\mu \beta} b_\nu b_\alpha \right). \ee We are now in position to discuss the result of equation (\ref{result}). First, as discussed in the last section, the constant $\alpha$ vanishes if a gauge invariant regularization technique is used, although the gauge symmetry of the model could be preserved if the Lorentz invariant and Lorentz-violating sectors are treated independently. However, the value of $\alpha$ is irrelevant here, since it can be absorbed in the other finite term, which depend on an arbitrary mass scale parameter. Second, the presence of a divergent term in the CPT-even coefficient is an important point to be analyzed. This indicates that the original classical action must contain such a term. In other words, the inclusion of a chiral nonminimal coupling in a modified QED requires the presence of the aether term from the beginning. This divergent factor multiplies also a Maxwell term. This means that the Lorentz preserving sector is also affected by the presence of the nonminimal interaction of chiral character. This is in contrast with the case of standard nonminimal coupling treated in papers \cite{mgomes1}, \cite{mgomes2} and \cite{scarp-NM}. In that case, the correction to the Maxwell term (and also the induced aether term) is finite and arbitrary, with the possibility of being set to zero. Last but not least, we must take into account that our model is nonrenormalizable. We have carried out an one-loop calculation and, at this order in the perturbative expansion, it has been shown that a new term which violates Lorentz symmetry and is CPT-even should be included in the classical action. If we go beyond the one-loop order, certainly new other terms will have to be considered. The nonrenormalizability of the model tells us that there is no a finite number of counterterms that will be sufficient to renormalize the theory. So, if we would like to deal with this effective model, we will have to stop at one-loop order. For this, it is necessary to find a cutoff energy $\Lambda$. This can be done like in the case of the simple nonminimal coupling, discussed bellow. We have to note that higher order terms in the coupling constant will allow for higher power contributions in the Lorentz violating parameter, with an increasing of the degree of divergence of the integrals. However, as demonstrated in \cite{Belich2} for the vectorial nonminimal coupling, the magnitude of the background vector is extremely small. We can impose the reasonable condition for the recovering of QED, $|b^2| \Lambda^2 << 1$. Since the effect of the divergences can be seen in a simplified form by substituting $m^2$ by $\Lambda^2$ in the coefficients, higher order calculations will furnish us higher powers of $|b^2| \Lambda^2$. So, although we can not prevent the proliferation of new terms beyond one-loop order, the predictability of such effective model is assured by the cutoff inequality above. This happens because at the same loop order each nonminimal vertex contributes with one factor of $b_\mu$, whereas higher loop orders with a fixed number of nonminimal vertices are controlled by the smallness of the coupling constant. In an effective model, the cutoff energy is an important parameter which should be established on physical grounds. Important features of a Quantum Field Theory, like causality and stability, can be lost at high energies \cite{causality1}, \cite{causality2}. The condition imposed by the inequality $|b^2| \Lambda^2 << 1$ is so that higher power terms in $b_\mu$ become less significant. In \cite{Belich2}, it has been established a bound such that $g \cdot |b_\mu|<10^{-32}\,(eV)^{-1}$ for the simple magnetic coupling. This bound and the inequality we propose above assure that such effective model is not considered at energies beyond the Planck scale. So, it is desirable that a calculation such that of \cite{Belich2} would be performed in order to establish a bound for the Lorentz violating parameter in the case of a chiral nonminimal coupling. In this case, we could consider the one-loop calculation meaningful. \section{Concluding Comments} An effective model for QED with the addition of a chiral nonminimal coupling has been investigated. This term, which is proportional to a fixed 4-vector $b_\mu$, violates Lorentz symmetry and originates a CPT-even Lorentz breaking term in the photon sector. Besides, since the coefficient of this quantum correction is divergent, such a model requires the presence of this aether term from the beginning in the classical action. This divergent factor multiplies also a Maxwell term. This means that the Lorentz preserving sector is also affected by the presence of the nonminimal interaction of chiral character. This is in contrast with the case of standard nonminimal coupling treated in the papers \cite{mgomes1}, \cite{mgomes2} and \cite{scarp-NM}. In that case, the correction to the Maxwell term is finite and arbitrary, with the possibility of being set to zero. The fact that the aether term should be present in the classical action from the beginning is problematic. If we go beyond one-loop order, the nonrenormalizability of the model tells us that there is no a finite number of counterterms that will be sufficient to renormalize the theory. So, if we would like to deal with this effective model, we will have to stop at one-loop order. For this, it is necessary to find a cutoff energy $\Lambda$. This can be done like in the case of the simple nonminimal coupling. In an effective model, the cutoff energy is an important parameter which should be established on physical grounds.Important features of a Quantum Field Theory, like causality and stability, can be lost at high energies \cite{causality1}, \cite{causality2}. The condition imposed by the inequality $|b^2| \Lambda^2 << 1$ is so that higher power terms in $b_\mu$ become less significant. In \cite{Belich2}, it has been established a bound such that $g \cdot |b_\mu|<10^{-32}\,(eV)^{-1}$ for the simple magnetic coupling. This bound and the inequality we propose above assure that such effective model is not considered at energies beyond the Planck scale. So, it is desirable that a calculation such that of \cite{Belich2} would be performed in order to establish a bound for the Lorentz violating parameter in the case of a chiral nonminimal coupling. In this case, we could consider the one-loop calculation meaningful. \subsection*{Acknowledgements} This work was partially supported by CNPq. The author wish to thank Marcos Sampaio and Prof. J. A. Helayel-Neto for illuminating discussions.
\section{Introduction} \label{sec:one} Massive star formation is thought to occur in clumps that are deeply embedded in the dense gas and dust of the natal molecular cloud. In these cold and dense environments, observations are often hindered by absorption or high optical depth of the emission and the confusion of many objects due to a limited angular resolution. Molecular transitions in the submm wavelength range are particularly useful probes to study these regions, since their critical densities are similar to the densities one expects for regions in which stars are forming. Even without resolving the star-forming clumps, one can find signposts of star formation, such as outflows and infall, from the kinematics derived from the shapes and radial velocities of molecular lines. In this paper, we try to use the molecular emission observed toward the sample of high extinction clouds, whose continuum and ammonia observations were discussed in \citet{rygl:2010b}, hereafter Paper I, to study the infall properties of the clumps and other indications for the onset of star formation. Crucial for the understanding of massive star formation is observational evidence for the infall of matter. In low-mass star-forming regions (SFRs) such as the Bok globule B335, infall has been inferred from the observed line profiles (\citealt{zhou:1993}). There are but few infall studies for candidate high-mass stars (\citealt{wu:2005,fuller:2005,beltran:2006a}); however, their number is currently increasing rapidly. A recent study by \citet{sepulcre:2010}, using a sample of infrared dark and infrared bright clumps from the literature, found a good correlation between the clump dust mass and the outflow mass, irrespective of the clump being infrared dark or bright. These authors claim that the outflow properties during star formation remain unchanged within a large range of masses and infrared properties of the clumps. For our infall study of dense clumps, we used not only the shapes of the $\mathrm{HCO^+}$(1--0) line but also the shapes of the $\mathrm{HCO^+}$(4--3), CO(3--2), and CO(1--0) emission lines. The different molecular transitions probe different densities and therefore different parts of the clump. Lower-$J$ CO lines trace the kinematics of the lower density material of the clumps. At high densities, $n\geq10^{4}\mathrm{cm^{-3}}$, and cold temperatures, $\sim$15\,K, the CO molecule depletes from the gas phase (\citealt{kramer:1999}). The level of CO depletion therefore indicates if the clump is cold and dense, and hence young, or if it consists of already more processed material. Apart from the presence of broad wings in lines from CO and other molecules, outflows can additionally be traced by an enhanced abundance of SiO, which is a good shock tracer (\citealt{schilke:1997}). We also observed $\mathrm{H_2CO}$\ and $\mathrm{CH_3CN}$ , which trace the more evolved stage of the young stellar object (YSO). Star formation is also evident by the presence of heating sources: a YSO's radiation increases the temperature of the dust and gas surrounding it. This is not only visible in infrared emission of the dust but also in several molecules like $\mathrm{H_2CO}$\ and $\mathrm{CH_3CN}$ . These molecules are often observed toward hot molecular cores, because the elevated temperatures ($>100$\,K) in these regions evaporate the icy grain mantles, increasing the gas phase abundances of such species by orders of magnitude (\citealt{mangum:1993, olmi:1996,tak:2000}). Additionally, for $\mathrm{CH_3CN}$\ one can estimate rotational temperatures of the hot phase from the emission in lines arising from different $K$ levels that cover a wide range in temperature, but are close together in frequency. Using these tracers, we define starless clumps as clumps without YSOs and signs of infall/outflows and prestellar clumps as clumps without YSOs but with infall/outflows. The higher density components of the clumps were studied with $\mathrm{N_2H^+}$\ and $\mathrm{H^{13}CO^+}$; $\mathrm{N_2H^+}$\ is known to be a reliable probe of cold gas with lower depletion than most other species (\citealt{tafalla:2002}). The hyperfine structure (hfs) of this molecule also allowed to estimate its excitation temperature and column density. The optically thin $\mathrm{H^{13}CO^+}$(1--0) line was used to determine the local standard of rest velocity ($v_\mathrm{LSR}$) of the clump. We tried to correlate star-formation behavior in the clump not only with surface density but also with the morphology and physical conditions of the cloud. In Paper I, we selected a sample of high extinction clouds based on our extinction maps from the first Galactic quadrant and performed 1.2\,mm dust continuum observations of them. We defined three categories of clouds based on their 1.2\,mm dust continuum contrast between the clump column density and the cloud column density. Clouds with a low clump to cloud column density contrast ($C_{N_\mathrm{H_2}}$) $C_{N_\mathrm{H_2}}$$<$2 were defined as {\em diffuse clouds} and thought to be the youngest clouds: they are the coldest and show few signs of star formation, such as masers and infrared emission. Clouds with higher contrast 2$<$$C_{N_\mathrm{H_2}}$$<$3 were defined as {\em peaked clouds} having at least one clump (or peak, hence `peaked' clouds) with a high contrast to the cloud emission; these clouds were thought to be the following stage. They show slightly higher temperatures and wider line widths (larger turbulence). The most evolved clouds were the {\em multiply peaked clouds}, defined by a still higher contrast $C_{N_\mathrm{H_2}}$$>$3 and generally containing several clumps (hence `multiply peaked' clouds). Most of the maser emission and infrared sources were found toward these clouds. Even though several clouds, mostly the peaked and multiply peaked ones, are elongated or filamentary, this phenomenological feature was not taken into account in defining the three cloud categories. The high extinction cloud sample that was studied in Paper I contained 50 clumps, of which 12 were in diffuse clouds, 19 in peaked clouds, and 19 in multiply peaked clouds. Almost all clumps showed $\mathrm{NH_3}$(1,1) emission, indicating the presence of high-density gas (Paper I). In the current paper, we report the results of a molecular survey on the clumps in high extinction clouds. With these results, we try to pinpoint their evolutionary stage and test our proposed evolutionary sequence of the clouds. We observed all the positions that contained $\mathrm{NH_3}$(1,1) emission with the IRAM 30m telescope in the $\mathrm{H^{13}CO^+}$(1--0), SiO(2--1), $\mathrm{HCO^+}$(1--0), $\mathrm{CH_3CN}$(5--4) $K$=0, 1, 2, 3, and 4 levels, $\mathrm{N_2H^+}$(1-0), $\mathrm{C^{18}O}$(2--1), and CO(2--1) molecular transitions. Next, a subsample of these sources was observed with the Atacama Pathfinder Experiment submillimeter telescope (APEX) in higher $J$ transitions: $\mathrm{N_2H^+}$(3--2), $\mathrm{H_2CO}$(4$_{03}$--$3_{04}$), $\mathrm{HCO^+}$(4--3), and CO(3--2). The APEX targets included mostly active and evolved sources, which are expected to emit strongly in these higher excitation transitions. In addition, a few diffuse and peaked clouds were added to allow comparison between the three categories of clouds. The line data were interpreted using the RADEX radiative transfer code (\citealt{tak:2007}) to arrive at models of the physical parameters of the clumps. The results of the line observations are compared with results from similar surveys toward other samples, e.g.\ the study of line emission toward Infrared Dark Clouds (IRDCs) and other massive star-forming regions by \citet{ragan:2006}, \citet{motte:2007}, \citet{pillai:2007}, \citet{pirogov:2007}, \citet{purcell:2009}, \citet{sepulcre:2010}, and \citet{vasyunina:2011}; the survey of methanol maser sources by \citet{purcell:2006}; the high-mass protostellar objects survey by \citet{fuller:2005} and \citet{beuther:2007b}; and similar work toward hot molecular cores by \citet{araya:2005} and UCH{\sc ii}\,regions by \citet{churchwell:1992}. This comparison will allow assessment of the differences in evolutionary stages covered by these studies. The observations and data reduction are described in Sect.~2. Sect.~3 gives the results of the infall study, the derived temperature and density estimates, the search for YSO indications, and the CO depletion study. These results are interpreted in light of an evolutionary sequence of the three classes of clouds in the discussion (Sect.~4) and summarized in Sect.~5. \section{Sample of high extinction clumps and observations} \subsection{The sample} The high extinction clouds were selected from color-excess maps in the 3.6\,$\mu$m--4.5\,$\mu$m {\it Spitzer} IRAC bands (\citealt{fazio:2004}). The maps cover Galactic longitudes $10^\circ<l<60^\circ$ in the first quadrant, $295^\circ<l<350^\circ$ in the fourth quadrant, and $-1^\circ<b<1^\circ$ in Galatic latitude. The method to construct the extinction maps is described in Paper I. We selected these clouds based on a 3.6\,$\mu$m--4.5\,$\mu$m color excess (CE) above 0.25\,mag, which corresponds to a visual extinction 20.45\,mag by $A_V=81.8\times CE(3.6\,\mu\mathrm{m}-4.5\,\mu\mathrm{m})$ using the reddening law of \citet{indebetouw:2005} (for more details, see Paper I). To guarantee visibility with the Effelsberg 100m and the IRAM 30m telescopes, we focused on the clouds in the first Galactic quadrant. In Paper I we analyzed the dust continuum emission and the ammonia inversion transitions of the clumps in high extinction clouds to derive their physical properties: the clumps have masses between 10 up 400\,$M_\odot$, they are cold, $\sim$16\,K, and are found at distances between 1 and 7\,kpc, with the majority being at 3\,kpc. A summary of the properties of the diffuse, peaked, and multiply peaked high extinction clouds, based on Paper I, is presented in Table \ref{ta:sample}. Additionally, we calculated the surface density, following \citet{sepulcre:2010} Eq. 1: $\Sigma = 4 M / \pi D^2$, where $M$ is the clump mass in grams and $D$ the diameter of the clump in cm. \begin{table*} \centering \caption{Properties of clumps in high extinction clouds based on Paper I\label{ta:sample}} \begin{tabular}{llccccccc} \noalign{\smallskip} \hline\hline \noalign{\smallskip} && \multicolumn{2}{c}{diffuse clouds}& \multicolumn{2}{c}{peaked clouds}&\multicolumn{2}{c}{multiply peaked clouds}& all clouds\\ && mean& \multicolumn{1}{c}{range}& mean& \multicolumn{1}{c}{range}& mean& \multicolumn{1}{c}{range}& mean\\ \noalign{\smallskip} \hline \noalign{\smallskip} cloud mass &$M_\odot$ &495 & 17 -- 3039 & 1420 & 70 -- 5500& 1900 & 500 -- 6500 & 910\\ cloud diameter &pc& 0.8 & 0.2 -- 2.3& 1.7 & 0.7 -- 3.7 & 1.9 & 1.2 -- 3.8& 1.2 \\ \noalign{\smallskip} clump mass &$M_\odot$&105 & 12 -- 283 & 130& 28 -- 738& 185 & 21 -- 431 & 150 \\ clump $N_\mathrm{H_2}$ &$10^{22}\,\mathrm{cm^{-2}}$& 5.4 & 2.9 -- 9.5 & 5.5 &3.3 -- 7.7& 8.5& 2.5 -- 26.3 &6.8\\ clump diameter &pc &0.3 & 0.11 -- 0.56 & 0.33 & 0.14 -- 0.76 &0.36 & 0.15 -- 0.47& 0.3\\ clump $\Delta v$ &$\mathrm{km~s^{-1}}$& 1.2 & 0.7 -- 2.3&1.4 & 0.9 -- 2.5 &1.6 & 0.8 -- 2.8 & 1.5\\ clump d &kpc&2.8 &1.1 -- 5.7& 3.4 & 1.9 -- 7.2& 3.6 & 2.1-- 4.7 & 3 \\ clump temperature & K &13.5 & 9.3 -- 16.7 & 15.7 & 11.9 -- 18.6 & 17.5 & 12.4 -- 24.7 & 16\\ clump $\Sigma$ &$\mathrm{g~cm^{-2}}$&0.3 &0.2 -- 0.5 & 0.3 & 0.2 -- 0.5& 0.5 & 0.2 -- 1.5&0.4\\ \noalign{\smallskip} \hline \hline \end{tabular} \tablefoot{Rows are (from top to bottom) cloud mass, cloud diameter, clump mass, clump column density, clump diameter, clump line width, clump kinematic distance, clump rotational $\mathrm{NH_3}$\ temperature, and clump surface density. } \end{table*} \subsection{IRAM 30m observations} The spectral line survey of the clumps in high extinction clouds was performed with the IRAM 30m telescope in 2007, June. We exploited the possibility of the ABCD receivers to observe simultaneously at two frequencies, one at $\sim$100\,GHz and the other at $\sim$230\,GHz. With two different receiver setups, we observed a total of seven molecular lines, listed in Table \ref{ta:freq}. The SiO(2--1) and $\mathrm{H^{13}CO^+}$(1--0) transitions were observed simultaneously in the same backend, since they are only separated by 100\,MHz. The higher frequency CO(2--1) and $\mathrm{C^{18}O}$(2--1) transitions were observed in both setups, which increased the signal-to-noise ratio by $\sqrt{2}$. For the $\sim$100\,GHz lines, we used several VESPA backend settings: one with a (relatively) narrow bandwidth of 40\,MHz and a channel spacing of 0.02\,MHz, corresponding to velocity resolution of 0.08\,km~s$^{-1}$, and two wider bandwidths of 120 and 160\,MHz using a channel spacing of 0.04\,MHz, corresponding to a velocity resolution of 0.16\,km~s$^{-1}$. The wider bandwidths were necessary to allow the SiO(2--1) and $\mathrm{H^{13}CO^+}$(1--0) lines to be observed together in the 160\,MHz bandwidth and to enable the broad, multiple-$K$ $\mathrm{CH_3CN}$(5--4) line profile to fit within the 120\,MHz band. For the $\sim$230\,GHz CO(2--1) and $\mathrm{C^{18}O}$(2--1) lines, which both have very wide line profiles, we used the 1~MHz backend, which offers a bandwidth of 512\,MHz and a spectral resolution of 1.5\,km~s$^{-1}$. Table \ref{ta:freq} gives an overview of the bandwidth and spectral resolution for each transition. We observed in position-switching mode, where the off position\footnote{The off position is an observation of `blank' sky, which is observed to remove the instrumental bandpass structure.} was located 800$''$\ away. During the observations, we performed a pointing check every 1.5 hours on a nearby quasar or on the H{\sc ii}\,regions G10.6--0.4 or G34.3+0.2. The pointing was found to be accurate within 4$''$. The focus check was usually performed on Jupiter or 3C\,273 at the beginning of each observing run. The opacity at 230\,GHz was variable from excellent winter weather conditions ($\tau=0.1$) to average summer conditions of $\tau=0.5$. The system temperatures at 230\,GHz ranged from 270--920\,K. At 100\,GHz the opacity ranged from 0.04--0.1, and the system temperatures were between 87 and 173\,K. The observed output counts were calibrated to antenna temperatures, $T^\star_{\mathrm{A}}$, using the standard chopper-wheel technique (\citealt{kutner:1981}) The $T^\star_{\mathrm{A}}$ temperature is the brightness temperature of an equivalent source that fills the entire $2\pi$ radians of the forward beam pattern of the telescope. This can be converted to a main beam brightness temperature, $T_{\mathrm{MB}}$, by multiplying by the ratio of the forward efficiency, $F_{\mathrm{eff}}$, and the main beam efficiency, $B_{\mathrm{eff}}$: \begin{equation} T_{\mathrm{MB}} = \frac{F_{\mathrm{eff}}}{B_{\mathrm{eff}}} \times T^\star_{\mathrm{A}}\,. \end{equation} Table \ref{ta:freq} lists the efficiencies for each transition. \subsection{APEX observations} The observations were carried out with the APEX 12m submillimeter telescope located on the Chajnantor plateau in the Atacama desert (Chile) during several runs between 2007 June 10 and November 2. We used the double sideband receiver APEX-2A equipped with two fast Fourier transform spectrometer (FFTS) backends (\citealt{risacher:2006,klein:2006}). The signal and image sidebands are separated by 12\,GHz. We used two setups in our observations, which are described in Table \ref{ta:freq}. Each FFTS has a bandwidth of 1 GHz, and 8192 channels, which corresponds to a velocity resolution of $\sim$0.12--0.15\,km~s$^{-1}$ for lines between 285--350\,GHz. For the $\mathrm{H_2CO}$(4--3) rotational transition, we observed only the $4_{04}-3_{03}$ $K$ level, since the remaining $K$ levels of this transition were outside the bandpass due to a mistake in the center frequency. Each observation was performed with on-off iterations using three subscans, integrating in total between 90--230 seconds on source. From the IRAM 30m observations, which were carried out before the start of our APEX runs, it was clear that the off positions were often contaminated by extended CO emission. Therefore, we used for the APEX observations off positions much further out, at 1800$''$ , which in most cases were free of emission. The calibration of the APEX data was, just as with the IRAM 30m telescope, carried out using the chopper-wheel technique. The telescope efficiencies, necessary to calculate the main beam brightness temperature, are listed in Table\.ref{ta:freq}. Before each run, the focus was adjusted on a planet, usually Jupiter. Pointings were made once every three hours on nearby H{\sc ii}\,regions with strong dust continuum emission, such as Sgr\,B2(N). The precipitable water vapor was between 1.44 and 2.80\,mm, while the system temperatures ranged from 190--300\,K at 280\,GHz and 410--575\,K for 356\,GHz. \subsection{Data reduction} The processing of the data, such as smoothing (spectral averaging) to increase the signal-to-noise ratio and baseline subtraction, was done in CLASS (\citealt{pety:2005}). For most lines, Gaussian fitting was performed to retrieve the basic line parameters, such as line widths, $v_\mathrm{LSR}$, and line intensities. For the lines with hyperfine structure (hfs), we used the hfs method, which allows additionally the derivation of the optical depth of the main hfs component (when applicable). The spectral plots were also prepared in CLASS. \begin{table*} \centering \caption{Molecular lines and frequencies\label{ta:freq}} \begin{tabular}{lcrrrrccccc} \noalign{\smallskip} \hline \hline \noalign{\smallskip} Molecule & $J+1\rightarrow J$ & $E_u/k_B$&Frequency & $n_\mathrm{crit}$\tablefootmark{a} & Bandwidth & Resolution &Beam &$B_\mathrm{eff}$&$F_\mathrm{eff}$& 1$\sigma$ r.m.s\\ & & (K)&(GHz) & (cm$^{-3}$)& (MHz) & (km~s$^{-1}$) & ($''$) & && (K) \\ \noalign{\smallskip} \hline \noalign{\smallskip} \multicolumn{10}{c}{IRAM 30m telescope}\\ \noalign{\smallskip} H$^{13}$CO$^+$ & 1--0 & 4.16~~ & 86.754~~ & $1.7\times10^5$ & 160 & 0.16& 29 &0.78 & 0.98 & 0.06 \\ SiO & 2--1 & 6.25~~ & 86.847~~ & $7.3\times10^5$ & 160 & 0.16& 29 &0.78 & 0.98 &0.07 \\ HCO$^+$ & 1--0 & 4.28~~ & 89.189~~ & $1.8\times10^5$ & 40 & 0.08 &28 & 0.78 & 0.98 &0.10 \\ CH$_3$CN & 5--4, $K$=0, 1, 2, 3 & 13.24\tablefootmark{b} & 91.987\tablefootmark{b} & $4.7\times10^5$ & 120 & 0.16&27 & 0.78 & 0.98 &0.09 \\ N$_2$H$^+$ & 1--0 & 4.47~~ & 93.174~~ & $1.6\times10^5$ & 40 & 0.08&27 & 0.78 & 0.98 &0.11 \\ C$^{18}$O & 2--1 & 15.81~~ & 219.560~~ & $9.2\times10^3$ & 512 & 1.5&12 & 0.62 & 0.94 &0.12 \\ CO & 2--1 & 16.60~~ & 230.538~~ & $1.1\times10^4$ & 512 & 1.5&11 & 0.58 & 0.92 &1.3 \\ \noalign{\smallskip} \multicolumn{10}{c}{APEX telescope}\\ \noalign{\smallskip} N$_2$H$^+$ & 3--2 & 26.83~~ & 279.512~~ &$3.0\times10^6$ & 1000& 0.15 & 22&0.73 &0.97&0.23 \\ H$_2$CO & $4_{04}$--$3_{03}$\tablefootmark{c} &34.90~~ & 290.623~~ &$9.2\times10^6$ & 1000& 0.15 & 22&0.73 &0.97&0.20 \\ $\mathrm{HCO^+}$\ & 4--3 & 42.80~~ & 356.734~~ &$9.1\times10^6$ & 1000& 0.12 & 18&0.73 &0.97&0.47\\ CO & 3--2 & 33.19~~ & 345.796~~ &$3.5\times10^4$ & 1000& 0.12 & 18&0.73 &0.97&1.1\\ \noalign{\smallskip} \hline \hline \end{tabular} \tablefoot{Columns are (from left to right) the molecule, its transition, its upper energy level, its frequency, its critical column density, the bandwidth used in the observation, the velocity resolution of the observation, the telescope beam, the main beam and forward beam efficiency of the observation, and the mean 1$\sigma$ noise value. \tablefoottext{a}{Calculated from the collision rates at $T=20$\,K from the LAMBDA molecular database (\citealt{schoier:2005}).} \tablefoottext{b}{Frequency and $E_u/k_B$ for the $K$=0 level. Observed were $K$=0, 1, 2, 3, 4} \tablefoottext{c}{The observed $K$ level is given in subscript.} } \end{table*} \section{Results} An overview of all the detected lines per clump and the clump J2000 positions are given in Table \ref{ta:det}. For sources with no detection, we put a 3$\sigma$ upper limit (also in Table~\ref{ta:det}). \subsection{Infall} \label{sect:infall} The infall signature of a source can be recognized by a line profile with a double-peaked structure, where the intensity of the blue peak exceeds the intensity of the red peak (\citealt{leung:1977,zhou:1993,myers:1996,mardones:1997, evans:1999}). Infall models assume a source where the infall velocity, density, and excitation temperature increase toward the center. In this scenario, the blue peak of an optically thick self-absorbed line originates in the rear part, and the red peak in the front part of the infalling shell. An increase in the infall velocity, $v_\mathrm{in}$, can cause the red peak to diminish; at very large values of $v_\mathrm{in}$, the red peak can even disappear and become a red ``shoulder''. A red excess, in contrast to this blue excess, may be caused by expansion or outflow. Alternatively, the red excess can be caused by an outwards moving blob of matter, instead of large-scale outward motion (\citealt{evans:1999}). To interpret the infall profile of an optically thick line, it is also necessary to compare it with an optically thin line measured toward the same position. The latter, often a line from a rare isotopologue, shows a maximum, defining the systemic velocity at the self-absorption minimum of the optically thick line. The observed line profiles can, in addition to large scale motions such as infall, also be influenced by abundance changes though the clouds. Hence, it is desired to investigate infall using several molecules. To probe different depths in the clouds and thereby possible different infall velocities, we used lines with different critical densities: the CO(2--1), CO(3--2), $\mathrm{HCO^+}$(1--0), and $\mathrm{HCO^+}$(4--3) transitions. \subsubsection{CO line profiles} We started the study with the CO(2--1) and CO(3--2) lines, whose critical densities are around 1$\times10^4$ and 4$\times10^4\,\mathrm{cm^{-3}}$ (Table \ref{ta:freq}), respectively. Since the CO molecule, as a common component of the interstellar medium (ISM); has a very extended distribution, the off position is not necessarily free of its emission. This can result in artificial absorption lines in the spectrum. To avoid this, one can observe with two different off positions and/or use different CO transitions. Since higher $J$ transitions have a higher critical density, their emission is confined to a more compact and higher density region than emission from lower $J$ lines. We employed this strategy and observed, in addition to the CO(2--1) line with the IRAM 30m, the CO(3--2) line with APEX using different off positions. Also, because of the molecules' ubiquity and relatively low critical density, CO observations pick up emission from various Galactic arms. This manifests itself in the spectrum by more than one peak at various local standard of rest (LSR) velocities. We searched for infall signatures by comparing the velocities of the CO(2--1) and CO(3--2) emission peaks with those of the optically thin $\mathrm{C^{18}O}$(2--1) line. When the peak of the optically thick line of the main isotopologue appeared shifted blueward of the $\mathrm{C^{18}O}$\ peak, the source was marked a blue excess source or a blue source. A source with a red shifted peak was marked a red source. In general, both the CO transitions showed similar behavior (see Table \ref{ta:infall}). Two examples of infall and outflow signatures are presented in Fig. \ref{fig:co}. When the peaks of the optically thick and thin lines were at the same velocity, nothing could be inferred. Mostly, the CO lines were broader than the $\mathrm{C^{18}O}$\ lines and had line wings extending from 10 to almost 50\,$\mathrm{km~s^{-1}}$ relative to the systemic LSR velocity. Line width broadening is partially due to a high optical depth, which is higher for the CO lines than for $\mathrm{C^{18}O}$\ lines. However, the wings are dominated by gas undergoing strong and dominant motions, such as infall or outflow, and in cases such as G022.06+00.21\,MM1 by high-velocity emission features (`bullets'). The results for CO are given in Table \ref{ta:infall}. For the lower transition, the presence of wings is indicated, and, if present, their velocity range is given. The CO line profiles delivered the first indications of infall or outflow. A deeper study was done with the $\mathrm{HCO^+}$(1--0) and $\mathrm{H^{13}CO^+}$(1--0) lines, reported in the next section. \begin{figure} \centering \includegraphics[angle=-90,width=6cm]{g24vfix.eps} \includegraphics[angle=-90,width=6cm]{g22-vfix.eps} \caption{Two examples of infall/outflow signatures of the $^{12}$CO line profiles. The optically thin $\mathrm{C^{18}O}$(2--1) emission (scaled for better visibility, shown in green) indicates the systemic velocity. The CO(2--1) and CO(3--2) emission is shown in black and red, respectively. The panels show infall ({\sl top }) and outflow ({\sl bottom}).\label{fig:co}} \end{figure} \subsubsection{The skewness parameter $\mathbf{\delta v}$} \begin{figure} \centering \includegraphics[angle=-90,width=6cm]{g1463_mm1_hcovfix.eps} \includegraphics[angle=-90,width=6cm]{g16_mm1_hcovfix.eps} \includegraphics[angle=-90,width=6cm]{g22_mm1_hcovfix.eps} \caption{Three examples of infall/outflow signatures of the $\mathrm{HCO^+}$(1--0) line profile. Red marks both $\mathrm{H^{13}CO^+}$(1--0), which is optically thin, and the systemic velocity of the dense gas. Compared to this the shift of the peak of the $\mathrm{HCO^+}$(1--0) emission, in black, becomes clear. The panels show infall ({\sl middle and bottom}), outflow ({\sl bottom right}), and a case of central self-absorption, where both peaks are equal ({\sl top}). The profiles show various degrees of self-absorption.\label{fig:hcopdv}} \end{figure} The optically thick $\mathrm{HCO^+}$(1--0) line profiles were used for an in-depth study of infall and outflow motions in the clumps. The systemic velocity of a clump was determined from the optically thin $\mathrm{H^{13}CO^+}$(1--0) line. Examples of $\mathrm{HCO^+}$(1--0) line profiles are shown in Fig.~\ref{fig:hcopdv}. For half of the clumps, the $\mathrm{HCO^+}$(1--0) line showed a double-peaked profile with a self-absorption dip. Here, the ratio of the two intensity peaks, $\frac{T_{\mathrm{blue}}}{T_{\mathrm{red}}}$, was used to check whether the source showed blue or red excess in the emission (ratio listed in Table \ref{ta:infall}). When the $\mathrm{HCO^+}$(1--0) line profile showed only one peak, we compared its velocity with the velocity of the optically thin $\mathrm{H^{13}CO^+}$(1--0) peak. The measure of this shift in velocity is called the skewness parameter. Following the method outlined in \citet{mardones:1997}, the skewness parameter $\delta v$, is defined as \begin{equation} \delta v = \frac{v_\mathrm{thick} - v_\mathrm{thin}}{\Delta v_\mathrm{thin}}\,, \end{equation} where $v_\mathrm{thick}$ and $v_\mathrm{thin}$ are the LSR velocities of the peaks of the optically thick and thin lines, respectively; $\Delta v_\mathrm{thin}$ is the line width of the optically thin line. The line width and LSR velocity of the $\mathrm{H^{13}CO^+}$(1--0) line were retrieved by Gaussian fits (Table \ref{ta:hcop}). For the optically thick $\mathrm{HCO^+}$(1--0) line, the profiles often showed non-Gaussian shapes; therefore, the position of the peak was determined by eye (also given in Table \ref{ta:hcop}). After \citet{mardones:1997}, we adopt the definition of significant blue and red excess, namely for the former if $\delta v \leq -0.25$ and the latter when $\delta v \geq 0.25$. The skewness parameter was determined for the $\mathrm{HCO^+}$(1--0), $\mathrm{HCO^+}$(4--3), and CO(3--2) lines (all listed in Table \ref{ta:infall}). For all molecules, the values ranged between --1.5 and 1.5, except for G14.39-00.75B where the $\delta v$ of the $\mathrm{HCO^+}$(1--0) line was --5.79. The latter is due to a very blue shifted $\mathrm{HCO^+}$(1--0) peak, most likely caused by a high self-absorption in this source. The clumps in high extinction clouds have skewness parameters similar to values in the literature: --0.5 for UCH{\sc ii}\,regions (\citealt{Wyrowski:2006}), --1.5 to 1 for HMPOs (\citealt{fuller:2005}), --1 to 2 for Extended Green Objects (EGOs, \citealt{chen:2009}), which are shocked regions, and --0.5 for the cluster-forming clump G24.4 (\citealt{wu:2005}). Cross correlations of the skewness between the three different molecules (Fig.~\ref{fig:infallcomp}, clumps from peaked and multiply peaked clouds are color coded in light of the evolutionary sequence discussion in Sect.~\ref{sec:evodis}) show that the $\mathrm{HCO^+}$(4--3) line is the least sensitive to infall and/or outflow, while the $\mathrm{HCO^+}$(1--0) seems the most sensitive. Most of the $\mathrm{HCO^+}$(4--3) lines' $\delta v$ values are $<|0.25|$, while those of the $\mathrm{HCO^+}$(1--0) and CO(3--2) lines exhibit a range of skewness. Also, \citet{fuller:2005} found that the transitions with higher critical densities show less infall. These observations suggest that the infall occurs predominantly in the low-density environment, which is in contrast with the increasing infall velocity toward the center that is expected in the collapsing clump model. The observations can possibly be explained by the different optical depths of the $\mathrm{HCO^+}$\ transitions, because an infall profile requires the transition to be sufficiently opaque. Depending on the density of the infall environment, the transitions with a high critical density might not be opaque enough to observe the skewness profile, while transitions of lower critical density will be already sufficiently optically thick to observe the infall profile. For the $\mathrm{HCO^+}$(1--0) and CO(3--2) lines, the skewness seems to be correlated, which makes the infall determination based on these two molecules more reliable. \begin{figure*} \centering \includegraphics[angle=-90, width=11cm]{infallcomp.eps} \caption{\label{fig:infallcomp} Skewness parameter of the $\mathrm{HCO^+}$(1--0) line versus that of the $\mathrm{HCO^+}$(4--3) and $^{12}$CO(3--2) lines. The dashed lines mark the boundary of significant excess at $\delta v >|0.25|$. The blue triangles represent clumps in peaked clouds, while the red squares represent sources in multiply peaked clouds.} \end{figure*} The excess parameter shows the average behavior of the blue excess over the red excess sources. It is defined as $E = (N_\mathrm{blue}-N_\mathrm{red})/N_{\mathrm{tot}}$, where $N_\mathrm{red}$ is the number of red excess sources, $N_\mathrm{blue}$ the total number of blue excess sources, and $N_\mathrm{tot}$ the total number of all sources (\citealt{mardones:1997}). \begin{table} \centering \caption{Distribution of the skewness parameter per molecule\label{ta:bintest}} \begin{tabular}{lrrrrrrr} \hline\hline \noalign{\smallskip} Transition & $N_\mathrm{blue}$ & $N_\mathrm{red}$ &$N_\mathrm{skew}$& $N_\mathrm{tot}$ & $E$ &$P$ & $<\delta v>$\\ \noalign{\smallskip} \hline \noalign{\smallskip} $\mathrm{HCO^+}$(1--0) & 24 & 14 &38 & 47 & 0.22 & 0.07 & $-0.26$\\ $\mathrm{HCO^+}$(4--3) & 5 & 4 &9 & 18 & 0.06 & 0.50 & $-0.11$\\ CO(2--1)\tablefootmark{a} & 24 & 11 &35 & 35 & 0.37 & 0.02 & -\\ CO(3--2) & 8 & 7 &15 & 16 & 0.06 & 0.50 & $-0.31$\\ \noalign{\smallskip} \hline \end{tabular} \tablefoot{Columns are (from left to right) number of blue excess sources, number of red excess sources, total number of sources with a significant skewness, total number of sources, the excess parameter, the probability of the distribution to arise by chance, and the average skewness.} \tablefoottext{a}{The numbers of the CO(2--1) transition serve only as an indication.} \end{table} Table \ref{ta:bintest} lists the excess parameter and average skewness parameter by transition. For each transition, we calculated, through the binomial test, the probability $P$ that the distribution between blue and red excess sources occurred by chance. The binomial distribution is defined as \begin{equation} P = \binom {n}{k} p^{k} (1-p)^{(n-k)}\,, \end{equation} where $n$ is the total number of trials, $k$ the number of successes, and $p$ the success probability. In our case, $n$ is the total number of blue and red excess sources, $k$ the number of blue excess sources, and the success probability $p$ equals $0.5$ if the blue and red excess sources are randomly distributed. We can then calculate the possibility that the distribution of the number of blue sources {\em equal or higher} than the number observed arises by chance by adding all possibilities $P$($n$, $k$, $p$) + $P$($n$, $k$+1, $p$) + .. until $k=n$. The small probability $P$ of 7\% for $\mathrm{HCO^+}$(1--0) line in Table \ref{ta:bintest} indicates that there are significantly more blue excess sources than expected for a random distribution. On the other hand, the higher CO(3--2) and $\mathrm{HCO^+}$(4--3) transitions have a probability of 50\%. Hence, the number of blue excess sources is likely by chance and not significant. The results for the CO(2--1) line, $P=0.02$, should be treated with care, because the same definitions of blue and red sources do not apply. Due to the possible confusion in the line profiles, the CO(2--1) was just classified based on the peak emission with respect to the $\mathrm{C^{18}O}$(2--1) line without the calculation of $\delta v$. Therefore, the numbers in Tables \ref{ta:infall} and \ref{ta:bintest} have to be taken just as indications. In conclusion, the $\mathrm{HCO^+}$(1--0) line shows a significant blue excess and is the best indicator of infall for our sample of clumps. The excess parameter of the $\mathrm{HCO^+}$(1--0) line, 0.22, is similar to values found in a previous survey by \citet{fuller:2005} of high-mass protostellar objects reaching excesses of 0.29 and 0.31. Studies of low-mass star formation also report similar numbers (\citealt{mardones:1997, evans:2003}). \subsection{Temperature and density estimations from ${N_2H^+}$(1--0) and (3--2) transitions} The $\mathrm{N_2H^+}$(1--0) and $\mathrm{N_2H^+}$(3--2) rotational transitions have hfs arising from the interaction between the molecular electric field gradient and the electric quadrupole moments of the two nitrogen nuclei (\citealt{caselli:1995}). The $\mathrm{N_2H^+}$(1--0) rotational transition is split into seven hyperfine components. The three main hyperfine groups were well resolved in our observations (see Fig. \ref{fig:n2hp}), and for a few sources we could fit all seven components. In the higher transition, we were barely able to even resolve the three main groups of hfs components (Fig. \ref{fig:n2hp}). \begin{figure} \centering \includegraphics[angle=-90,width=6cm]{g1391_mm1_n2hpvfix.eps} \includegraphics[angle=-90,width=6cm]{g17_mm1_n2hp32vfix.eps} \caption{\label{fig:n2hp} $\mathrm{N_2H^+}$(1--0) ({\em top}) and $\mathrm{N_2H^+}$(3--2) ({\em bottom}) line profiles. The green lines show the locations of the hyperfine components.} \end{figure} For fitting the $\mathrm{N_2H^+}$(1--0) transition we took the hfs into account by using the hfs method for the line fitting in CLASS \footnote{http://www.iram.es/IRAMES/otherDocuments/postscripts/classHFS.ps}. This method assumes one excitation temperature and one line width for all seven hyperfine components, as well as the fact that the opacity as a function of frequency for each hfs has a gaussian shape. Then, besides the intrinsic line width and integrated intensities, the hfs method also determines the optical depth of the main hfs group. The total optical depth $\tau_\mathrm{tot}$ can be calculated from the obtained main hfs group optical depth $\tau_\mathrm{main}$ by multiplying it by the sum of relative intensities of the satellites $S$ as $\tau_\mathrm{tot}=S\tau_\mathrm{main}$. In case the relative intensities are scaled so that their sum equals unity, the hfs fit directly delivers the total optical depth (the sum of all the hfs optical depths). We calculated the column density of $\mathrm{N_2H^+}$, $N_\mathrm{N_2H^+}$ (\ref{eq:n2hp}) following \citet{benson:1998}. The results of the $\mathrm{N_2H^+}$(1--0) line profile fitting; the integrated intensity (including the hyperfine components), total optical depth, and line width are placed together with the excitation temperature and the column density in Table \ref{ta:n2hp10}. For $\mathrm{N_2H^+}$(3--2), Table \ref{ta:apex1} lists the parameters obtained by Gaussian fits. We did not observe the hfs as clearly as in the lower transition and could therefore not determine the optical depth for the higher transition. For sources with $\tau_\mathrm{(1-0),\,tot}\lesssim0.4$, which usually also had high relative errors, we found unrealistically high $T_\mathrm{ex}$ values that were much higher than the ammonia rotational temperatures, $T_\mathrm{rot}$ (Paper I). The CLASS hfs method does not provide good estimates for excitation temperatures (hence also column densities) when the relative errors in optical depth are very large ($>40$\%) and when the optical depth is very small ($\tau_\mathrm{(1-0),\,tot}\lesssim0.4$) (see, e.g., \citealt{fontani:2006,fontani:2011}). After discarding these sources, the excitation temperatures generally ranged from 3 to 29\,K, averaging 8.1\,K . In fact, only two sources had a $T_\mathrm{ex}>20\,K$; G017.19+00.81 MM2 (28\,K) and MM3 (29\,K). We assumed that the $\mathrm{N_2H^+}$\ gas has a similar kinetic temperature as the $\mathrm{NH_3}$\ gas and that they are both in local thermal equilibrium (LTE), in which case the $T_\mathrm{ex}$ should equal $T_\mathrm{rot}$. However, the determined excitation temperatures were, with the exception of G017.19+00.81 MM2 and MM3, all lower than the rotational temperature from ammonia. The difference between the $T_\mathrm{ex}$ determined here and the $T_\mathrm{rot}$ determined by ammonia is the $\mathrm{N_2H^+}$\ filling factor. While $T_\mathrm{ex}$, is inversely proportional to the filling factor (see \ref{app:n2hp}), the rotational temperature is an intrinsic temperature (independent of $f$ since it is solely determined by line ratios). We estimated the filling factor from the ratio of the excitation temperature of $\mathrm{N_2H^+}$(1--0) and the $\mathrm{NH_3}$\ rotational temperature: \begin{equation} f_{\mathrm{N_2H^+(1-0)}}\sim\frac{T_\mathrm{ex,\,N_2H^+}}{T_\mathrm{rot,\,NH_3}}. \end{equation} The individual filling factors range from 0.2 to 0.8. For clumps G17.19+00.81 MM2, and MM3, we found meaningless filling factors larger than unity, which is possibly due to an overestimation of the excitation temperature. A filling factor smaller than unity indicates either that the source size is smaller than the beam or that the source is clumpy instead of centrally condensed. Alternatively, the $\mathrm{N_2H^+}$(1--0) line could be sub-thermally excited, a consequence of densities lower than the critical density. At such densities, a transition's upper energy level becomes underpopulated: collisions do not manage to populate it, and the system is in non-LTE. In this case, the excitation temperature will be lower than the kinetic temperature, thus mimicking the behavior of a small filling factor. To have another estimate of the filling factor, we performed radiative transfer modeling to obtain the intrinsic $\mathrm{N_2H^+}$\ column density. In a non-LTE system, the ratios of line intensities deviate from the LTE predictions and become a function of density, column density, temperature, and optical depth. In particular, we consider the $\mathrm{N_2H^+}$(1--0)/$\mathrm{N_2H^+}$(3--2) line ratio and combine this with non-LTE radiative transfer modeling using RADEX (\citealt{tak:2007}) and the molecular database LAMBDA (\citealt{schoier:2005}). Figure \ref{fig:n2hp-ratio} shows the integrated intensities of the two $\mathrm{N_2H^+}$\ transitions plotted against each other. We use differently colored symbols for each cloud class (diffuse, peaked, and multiply peaked) in the graphs shown throughout this paper to visualize the (possible) behavior for different phases of cloud evolution. The integrated intensity, $\int T_{\mathrm{MB}}\mathrm{d}v$, of the lower transition, $\mathrm{N_2H^+}$(1--0), exceeded that of $\mathrm{N_2H^+}$(3--2) by an average of \begin{equation} <\frac{\int T_{\mathrm{MB}}\mathrm{d}v(\mathrm{N_2H^+}(1-0))}{\int T_{\mathrm{MB}}\mathrm{d}v(\mathrm{N_2H^+}(3-2))}>=2.9.\\ \end{equation} In this ratio, we assumed that the filling factor ratio is 1, since it is likely that the filling factors of the two transitions cancel out as the APEX beam is only slight smaller than the IRAM 30m beam (Table \ref{ta:freq} lists all beamsizes). There are three sources that lie far from the average: G014.63--00.57 MM1, G017.19+00.81 MM2, and G022.06+00.21 MM1, which are all bright at 24\,$\mu$m and contain water masers (see Table \ref{ta:det}). \begin{figure}[!h] \centering \includegraphics[angle=-90,width=7cm]{ratio-n2hp-class.eps} \caption{\label{fig:n2hp-ratio} $\mathrm{N_2H^+}$(1--0) integrated intensity versus $\mathrm{N_2H^+}$(3--2) integrated intensity for all detected sources. The blue triangles represent clumps in peaked clouds, while the red squares represent sources in multiply peaked clouds. The green solid line marks the trend of the observed ratio of 2.9.} \end{figure} \begin{figure*}[!h] \centering \includegraphics[angle=-90,width=13cm]{radexplot_newv2.eps} \caption{\label{fig:Nn-n2hp} $\mathrm{N_2H^+}$(1--0) optical depth versus the ratio of $\mathrm{N_2H^+}$(1--0) and $\mathrm{N_2H^+}$(3--2) integrated intensities. The four plots represent RADEX calculations with different temperatures: top left 10\,K; top right 15\,K; bottom left 20\,K; bottom right 25\,K. The black dashed contours represent the logarithm of the line width times the column density, $\log(\mathrm{N_{N_2H^+}}/ \Delta v )$ with a contour step of 0.1. The red dashed contours are $\log(n_{\mathrm{H_2}})$ with a contour step of 0.2. The lowest and highest contour values are given in each plot. The values of the sources in Table~\ref{ta:RADEX} are plotted by black dots.} \end{figure*} The observed $\mathrm{N_2H^+}$\ column density is roughly proportional (depending on the value of $T_\mathrm{ex}$, see Eq.~\ref{eq:n2hp}) with excitation temperature, $N_\mathrm{N_2H^+} \propto T_\mathrm{ex}^{1.4}$ for 5\,K$<T_\mathrm{ex}<$10\,K, which in turn is inversely proportional with filling factor. Hence, the ratio of observed $\mathrm{N_2H^+}$\ column density (from the $\mathrm{N_2H^+}$(1--0) line) and the intrinsic column density calculated by RADEX (based on the $\mathrm{N_2H^+}$(1--0)/$\mathrm{N_2H^+}$(3--2) integrated intensity ratio) can be interpreted as a result of the filling factor of the $\mathrm{N_2H^+}$(1--0) line. Using RADEX, we estimated the behavior of $N_\mathrm{N_2H^+}/\Delta v$ and $n_\mathrm{H_2}$ for different values of the $\mathrm{N_2H^+}$(1--0)/$\mathrm{N_2H^+}$(3--2) line ratio and the opacity of $\mathrm{N_2H^+}$(1--0) for kinetic temperatures of 10, 15, 20, 25\,K. We chose this temperature range based on rotational temperatures derived from the ammonia observations (see also Table \ref{ta:sample}). The results are shown in Fig.~\ref{fig:Nn-n2hp} and Table \ref{ta:RADEX}. \begin{table*} \centering \caption{Results of RADEX calculations based on the (1--0)/(3--2) $\mathrm{N_2H^+}$\ ratio \label{ta:RADEX}} \begin{tabular}{llccccccccc} \hline \noalign{\smallskip} &&\multicolumn{3}{c}{\underline{\hspace{1.2cm}$\mathrm{N_2H^+}$(1--0)\hspace{1.2cm}}} & \underline{\hspace{0.0cm}$\mathrm{N_2H^+}$(3--2)\hspace{0.0cm}}& &&&&\\ Source name& &$\int T_{\mathrm{MB}}\mathrm{d}v$&$\Delta v$ &$\tau$&$\int T_{\mathrm{MB}}\mathrm{d}v$&$T_\mathrm{rot}$ &$N_{\mathrm{N_2H^+, obs}}$ &$N_{\mathrm{N_2H^+, R}}$&$n_\mathrm{N_{H_2},obs}$&$n_\mathrm{N_{H_2},R}$\\ & & (K~km~s$^{-1}$) & (km~s$^{-1}$) & & (K~km~s$^{-1}$) & (K) & ($10^{13}\mathrm{cm^{-2}}$) & ($10^{13}\mathrm{cm^{-2}}$) &($10^5\mathrm{cm^{-3}}$) &($10^5\mathrm{cm^{-3}}$)\\ \noalign{\smallskip} \hline \noalign{\smallskip} G013.91--00.51&MM1 & 12.9 & 1.19 &5.34& 2.4 & 14.1 & 2.2& 5.8 & 2.2& 0.3\\ G014.63--00.57& MM1 & 45.8 &2.46& 2.35 & 25.6 & 18.1 & 8.9 &18.7 & 7.7&3.4\\ G014.63--00.57& MM2 & 17.9 & 1.27 &3.90 & 8.2 & 15.7 & 2.8 & 9.4 & 7.3 & 3.4\\ G017.19+00.81&MM2 &32.9 & 2.21 &1.0 &18.7 &18.7 & 13.5 & 7.7 & 6.3 & 4.0\\ G017.19+00.81&MM3 &25.1 & 1.74 &0.91 & 1.7 & 20.1 & 10.9 &2.7 & 0.8 & 1.8\\ G018.26--00.24& MM1&23.5 & 2.50 & 2.10 & 8.6 & 18.2 & 3.7 & 10.5 & 1.8 & 1.4\\ G018.26--00.24& MM2&26.0 & 2.16& 4.00 & 7.6 & 17.4 & 4.3 & 10.1 & 1.3 & 1.0\\ G022.06+00.21&MM1&28.7 & 2.50 &1.17 & 16.3 & 24.7 & 7.7 & 13.0 & 6.8 & 5.2\\ G024.37--00.15& MM2 & 12.7 & 2.28 &1.55 &5.4 & 15.7 & 2.1 & 6.2 & 1.2 & 2.9\\ G024.94-00.15&MM1 & 20.0 & 2.48 &1.30 & 7.4 & 15.2 & 4.2 & 5.0 & 2.0 & 2.9\\ G030.90+00.00A&MM1 & 22.8 &2.38 & 0.26 & 7.7 & 18.6 & -- & 3.1 & 0.7 & 1.8\\ G034.71--00.63&MM1 & 11.7 & 2.31 & 2.61&6.1 & 17.8 & 1.9 & 17.6 & 1.0 & 3.4\\ G034.71--00.63&MM2 & 11.5 & 1.58 & 4.90& 2.7 & 12.4 &2.1 & 6.3 & 1.1 & 1.4\\ G053.81--00.00 & MM1 & 9.0 & 1.64& 0.56&3.7 & 12.4 & 1.0 & 12.1 & 3.6 & 8.6\\ \noalign{\smallskip} \hline \end{tabular} \tablefoot{Column are (from left to right) source name, integrated $\mathrm{N_2H^+}$(1--0) line intensity, FWHP $\mathrm{N_2H^+}$(1--0) line width, $\mathrm{N_2H^+}$(1--0) optical depth, integrated $\mathrm{N_2H^+}$(3--2) line intensity, $\mathrm{NH_3}$\ rotational temperature (Paper I), observed $\mathrm{N_2H^+}$\ column density, calculated $\mathrm{N_2H^+}$\ column density from RADEX, observed hydrogen density from the 1.2\,mm continuum (Paper I), calculated hydrogen density from RADEX. } \end{table*} In Fig.~\ref{fig:Nn-n2hp} we marked the positions of the sources from Table \ref{ta:RADEX} at the intersection of the ratio of the (1--0) and (3--2) $\mathrm{N_2H^+}$\ intensities and the $\mathrm{N_2H^+}$(1--0) optical depth in the plot calculated with a kinetic temperature closest to the $\mathrm{NH_3}$\ rotational temperature of the clump. To obtain the intrinsic $\mathrm{N_2H^+}$\ column density, we multiplied $N_\mathrm{N_2H^+}/ \Delta v$ by the observed $\mathrm{N_2H^+}$(1-0) line width. In the plots, we see that a rising $T_\mathrm{kin}$ increases the column density for a given optical depth and a line ratio, while the volume density decreases. For the handful of sources that we have measured both $\mathrm{N_2H^+}$\ transitions, we compared the filling factor based on the column density with the one determined from the temperature. On average, the filling factors agreed within 30\%. Here we excluded clumps G017.19+00.81 MM2 and MM3, since column density-derived filling factor came out larger than unity, and which confirms that for these sources the derivation of the $\mathrm{N_2H^+}$\ parameters, from which the excitation temperature and column density are calculated, gave unrealistic values. The RADEX results for the volume density were almost all within a factor two or less of the measured volume density from the 1.2\,mm continuum (Table \ref{ta:RADEX}). Given the uncertainties in the derivations of both volume densities, we can say that they are in agreement for most of the sources. The $\mathrm{N_2H^+}$\ ratio and the $\mathrm{N_2H^+}$(1--0) optical depth can therefore be used to estimate the hydrogen volume density and intrinsic $\mathrm{N_2H^+}$\ column density of a clump for a given kinetic temperature. We estimated the abundance, $\chi$, of $\mathrm{N_2H^+}$\ by comparing the $N_\mathrm{N_2H^+}$ corrected by the filling factor from the temperature ratio (for each source individually) to the hydrogen column density, $N_\mathrm{H_2}$, measured from the 1.2\,mm continuum (Paper I): \begin{equation} \chi(\mathrm{N_2H^+})=\frac{N_\mathrm{N_2H^+}}{N_\mathrm{H_2}}. \end{equation} The individual clump abundances are listed in Table \ref{ta:n2hp10}, and range from $0.26-4.0 \times10^{-9}$, with a mean of $1.3\times10^{-9}$. The true abundances are likely to be slightly lower, since the $N_\mathrm{H_2}$ is a beam-averaged column density and contains an unknown filling factor. Taking this into account, the abundance is in agreement with the value of \citet{pirogov:2007} for high-mass star-forming cores and the results of \citet{ragan:2006, vasyunina:2011} for IRDCs. \subsection{Presence of young stellar objects} \subsubsection{Hot molecular cores: $\mathrm{CH_3CN}$ and $\mathrm{H_2CO}$ } \label{sect:myso} $\mathrm{CH_3CN}$\ and $\mathrm{H_2CO}$\ are usually found toward YSOs or hot molecular cores (\citealt{mangum:1993,olmi:1996,tak:2000}), which are warm ($\geq100$\, K) and dense ($\geq10^5-10^6$\,cm$^{-3}$) (\citealt{kurtz:2000}). Both molecules can be used as a tracer of ongoing star formation. $\mathrm{CH_3CN}$\ is a symmetric-top molecule where each rotational level is split for different projections of angular momentum along the symmetry axis of the molecule labeled by the quantum number $K$. Within one rotational level, $J+1 \rightarrow J$, the transitions of the $K$ components are determined solely by collisions, and their relative intensities are therefore related to the kinetic temperature of the region (\citealt{solomon:1971}). The $K$ components with one rotational level are very closely spaced in frequency and can be observed simultaneously with one backend with a bandwidth $\gtrsim$30\,MHz. Under the assumption that the $\mathrm{CH_3CN}$(5--4) $K$=0, 1, 2, 3, and 4 emission is optically thin and uniformly fills the beam and that all $K$ transitions within this $J$ level are characterized by one temperature, we derived the rotational temperatures, $T_\mathrm{rot}$, and the total $\mathrm{CH_3CN}$\ column density following \citet{araya:2005}. These authors studied the $\mathrm{CH_3CN}$\ emission toward objects similar to ours (massive star-forming clumps) and showed that it is reasonable to assume the optically thin limit for this line. A beam-filling factor smaller than unity would affect the $K$ levels similarly and hence would have no large influence on the temperature derivation if the filling factor does not vary from one $K$ component to another. The $\mathrm{CH_3CN}$\ column density, however, is affected by the beam-filling factor. Given the beam size of 27$''$\ against the average angular clump size found at 1.3\,mm, $\sim$20$''$, the beam-filling factor could be slightly smaller than unity. We detected the $K$=0 to 3 levels, or a subset of these, for several sources; the data were too noisy (1\,$\sigma$ rms of 0.9\,K) for a $K$=4 level detection. All detections, derived $T_\mathrm{rot}$, and column densities are given in Table \ref{ta:mecn}. The $\mathrm{CH_3CN}$\ rotation temperatures were much higher than the $\mathrm{NH_3}$\ temperatures (see Table \ref{ta:mecn} and Fig.~\ref{fig:trotch3cn}), since the $\mathrm{CH_3CN}$\ traces warmer gas than ammonia does. For the $\mathrm{H_2CO}$(4$_{03}$--3$_{04}$) line, the observed main beam brightness temperatures, LSR velocities, and line widths were determined from Gaussian fits and are listed in Table \ref{ta:sf}. \begin{figure} \begin{center} \includegraphics[angle=-90,width=9cm]{trot-8aug.eps} \caption{$\mathrm{CH_3CN}$\ rotational temperatures versus $\mathrm{NH_3}$\ rotational temperatures. Clumps for which the $\mathrm{CH_3CN}$\ temperature had errors smaller than 12\,K are red. \label{fig:trotch3cn}} \end{center} \end{figure} To understand if YSOs are preferably present in denser or more compact clumps, we plotted all the clumps with evidence for an embedded YSO in a size versus the column density graph (Fig.~\ref{fig:sftracer}). In addition to plotting clumps with evidence of a hot molecular core through the detection of $\mathrm{H_2CO}$\ and $\mathrm{CH_3CN}$ , we plot clumps with a 24\,$\mu$m detection (direct evidence of dust heated by an YSO) and SiO emission (evidence of outflows, see Sect.~\ref{sect:sio}). Apparently, YSOs are found in clumps of all sizes, but, more interestingly, they are mostly found at high column densities. The minimal $N_\mathrm{H_2}$ at which the $\mathrm{H_2CO}$(4${03}$--3$_{04}$) line was detected is $4\times10^{22}\,\mathrm{cm^{-2}}$, while the $\mathrm{CH_3CN}$(5--4) $K=0$emission was still observed toward one less dense clump with $N_\mathrm{H_2}=3\times10^{22}\,\mathrm{cm^{-2}}$. The column density cut-off is most clearly seen in the 24\,$\mu$m and SiO line detections; above a column density of $5\times10^{22}\,\mathrm{cm^{-2}}$, almost all clumps show infrared emission and signs of outflows. For both the $\mathrm{CH_3CN}$\ and $\mathrm{H_2CO}$\ molecules, there is a general cut-off in the column density at $4\times10^{22}\,\mathrm{cm^{-2}}$, below which no YSOs are found. This cut-off is slightly lower than the column density of $\sim$$1\times10^{23}\,\mathrm{cm^{-2}}$ theoretically required to form massive stars (0.7 g\,cm$^{-2}$, \citealt{krumholz:2008}). \citet{sepulcre:2010} found a column density threshold of $7.7\times10^{22}\,\mathrm{cm^{-2}}$, below which the outflow rate decreases rapidly and the outflows are less massive. Hence, while observational data seem to strongly support a column density cutoff for (massive) star formation, they also seem to indicate a slightly lower threshold than theoretically advocated. The difference, however, is not too large, and the observed column densities are beamaverages, which means that if the source size is smaller, we are underestimating the true column density. \begin{figure}[!h] \centering \includegraphics[width=9cm,angle=-90]{mysosio-8aug.eps} \caption{\label{fig:sftracer} High extinction clumps placed in a diagram of column density versus the FWHM diameter (in pc) derived from the 1.2\,mm continuum (Paper I). The detections of MIPS 24\,$\mu$m emission, $\mathrm{H_2CO}$(4$_{03}$--3$_{04}$), $\mathrm{CH_3CN}$(5--4) $K$=0, and SiO(2--1) are marked in color, black symbols are non-detections. The crosses mark the clumps in diffuse clouds, triangles clumps in peaked clouds, and squares clumps in multiply peaked clouds.} \end{figure} \subsubsection{Outflow: SiO(2--1) emission} \label{sect:sio} Apart from evidently heating their environments, (massive) YSOs drive strong molecular outflows (\citealt{beuther:2002b}), which may remove, in the case of disk accretion, the angular momentum away from the forming star. While mechanisms to accelerate the outflow have been proposed for low-mass star formation (\citealt{shu:2000}), in the high-mass scenario the acceleration mechanism is not well understood yet. Interstellar dust grains contain silicates (\citealt{draine:2003}). Shock waves driven by outflows can sputter Si-bearing compounds in dust grains leading to increased SiO abundances in the post-shock gas (\citealt{schilke:1997}). Thus, SiO is expected to be associated with ongoing star formation via its formation history. Observations support this, since SiO is more common in more active (or evolved) sources, while very weak in quiescent (hence early) sources (\citealt{beuther:2007b}). Naturally, one also expects to observe wide line widths. We found line widths from $\sim$1.5--7\,km~s$^{-1}$ for very weak SiO detections, while for stronger SiO emitting sources the line widths increase up to 26\,km~s$^{-1}$. The individual values for all detections are given in Table \ref{ta:sf} along with the main beam brightness temperatures and LSR velocities. The mean noise level of the SiO observations is 0.07\,K, which is comparable to previous studies in massive star-forming regions by \citet{beuther:2007b} and \citet{motte:2007}, which reached noise levels of 0.02\,K and 0.10\,K, respectively. \begin{figure} \centering \includegraphics[angle=-90,width=7cm]{sio-infall-classv2.eps} \caption{\label{fig:sio-infall} SiO detections against the $\mathrm{HCO^+}$(1--0) skewness parameter $\delta v$. The green crosses represent sources in diffuse clouds, the blue triangles clumps in peaked clouds, and the red squares sources in multiply peaked clouds.} \end{figure} We detected the SiO(2--1) transition toward 50\% of the whole sample. Of all the SiO-emitting sources, a remarkable majority, 70\%, coincided with infall sources (having a $\mathrm{HCO^+}$(1--0) skewness parameter, $\delta v\leq -0.25$), as illustrated in Fig.~\ref{fig:sio-infall}. This finding indicates that accretion is accompanied by collimated outflows, which cause the enhancement of SiO. A further 15\% of the SiO detections are sources exhibiting expansion ($\delta v\geq 0.25$), and the remaining 15\% are sources showing neither expansion nor infall ($-0.25 < \delta v < 0.25$). Clumps without SiO detections have $\mathrm{HCO^+}$(1--0) skewness parameters divided almost equally between infall (ten clumps) and expansion (eight clumps). \begin{figure} \centering \includegraphics[angle=-90,width=7cm]{sio-dv-ncol.eps} \caption{\label{fig:sio} SiO line width against the hydrogen column density $N_\mathrm{H_2}$. The green crosses represent sources in diffuse clouds, the blue triangles clumps in peaked clouds, and the red squares sources in multiply peaked clouds} \end{figure} To check if the SiO line widths, and hence the outflows, are stronger for the more evolved sources, we compared them with the hydrogen column density determined from the 1.2\,mm continuum. Figure~\ref{fig:sio} shows that up to column densities of $10^{23}\,\mathrm{cm^{-2}}$ there is a tendency of increasing line widths with column density. The few clumps that are above this column density limit, G014.63--00.57 MM1, G017.19+00.81 MM2, G022.06+00.21 MM1, and G014.63--00.57 MM2, show no clear trend. The SiO emission was detected toward all multiply peaked clouds, but just toward three peaked clouds (25\%) and only two diffuse clouds (18\%), singling out the multiply peaked clouds as active regions of star formation. \subsubsection{Prestellar and starless sources} Our sample contained 19 sources with no evidence of a YSO. Of these sources, one-third shows infall in the $\mathrm{HCO^+}$(1--0) line profile indicating that these are in a {\em prestellar phase}, one-third is expanding according to the $\mathrm{HCO^+}$(1--0) skewness parameter, and the last third shows neither infall nor outflow. The sources without any signs of infall or YSOs were defined as {\em starless}. \citet{motte:2007} showed how difficult it is to find high-mass analogies of low-mass starless cores, such as L1521F (\citealt{shinnaga:2004}), in their study of the Cygnus X complex. Remarkably, they found SiO emission in all of the observed infrared-quiet massive dense cores, indicating that star formation is already ongoing. In our sample, we found six prestellar and six starless sources. Only half of them contain a clump (an overdensity of typical hydrogen column density of $10^{22}\,\mathrm{cm^{-2}}$, which could be identified by the Miriad routine ``sfind", see Paper I); most of them are too diffuse in the mm continuum. The masses of the prestellar clumps, derived from the 1.2\,mm continuum, were greater than their virial masses, indicating that gravity is the dominant force. However, also for the starless clumps, their virial parameters indicate that gravity dominates their internal dynamics. Apparently, it is not merely the lack of gravitational force that separates the starless from the prestellar clumps. The prestellar and starless sources were found in the diffuse and peaked clouds, but not in the multiply peaked clouds, which agrees with the previous indications of the multiply peaked clouds as the most evolved clouds. Most of the starless sources were in diffuse clouds, which are dark in the 24\,$\mu$m MIPS images. The prestellar sources were generally found in the vicinity of infrared objects, most likely more evolved YSOs. These YSOs are possibly responsible for introducing shocks into their surrounding medium through stellar winds and outflows, thus supporting the formation of higher densities regions where gravitational collapse can occur. This implies that stars form close to other stars, which is a well-known fact and agrees with the observed clustering of newly formed stars (see, for example, \citealt{lada:2003}). The limited number of both starless and prestellar sources in a total of 12 clumps, compared to the 31 sources with a YSO indication, agrees with the findings of \citet{motte:2007} that the starless and prestellar phases are short and dynamic. \subsection{CO depletion} As a clump contracts, the level of depletion of C-bearing molecules like CO or CS increases (\citealt{kramer:1999,bergin:1997}). The CO depletion in a clump can therefore be used as a time marker. Observations of massive protostellar objects (\citealt{purcell:2009}) and clumps in IRDCs (\citealt{pillai:2007}) show depletion of CO, indicating that these objects are already in a state of collapse and not in the chemically earliest stage. CO depletion as a time marker should be treated with some caution, because heating of the gas (by a YSO) can cause the CO to be desorbed to the gas phase, creating low CO-depletion levels as in the chemically earliest stages around the onset of collapse. We used $\mathrm{C^{18}O}$, a less common isotopologue of CO, to measure the depletion in the clumps of the high extinction clouds. In contrast to CO, the line profile of $\mathrm{C^{18}O}$ lines consisted of one Gaussian without any wings or self-absorption. This is an indication that the $\mathrm{C^{18}O}$(2--1) emission has a low or moderate optical depth, which would agree with previous observations that derived the $\mathrm{C^{18}O}$\ optical depths by comparing the $\mathrm{C^{18}O}$\ data with emission from the even rarer $\mathrm{C^{17}O}$ (\citealt{crapsi:2005,pillai:2007}). Therefore, in this work, we will assume that the $\mathrm{C^{18}O}$(2--1) line is optically thin. The $\mathrm{C^{18}O}$(2--1) line profiles were fitted with a Gaussian. Table \ref{ta:c18o-ii} lists our obtained integrated intensities, the LSR velocities, and the line widths. The column density of $\mathrm{C^{18}O}$ was calculated in the optically thin limit (see Eq.\ref{eq:c18o}) under the assumption of a beam filling factor of unity, which is likely for a beam of 11$''$\ and clumps that have an average size of 20$''$\ in the 1.2\,mm continuum. Second, we assume LTE and equate the excitation temperature to the rotational temperature, $T_\mathrm{rot}$, derived from the $\mathrm{NH_3}$(1,1) and $\mathrm{NH_3}$(2,2) transitions. In very dense and cold regions, where $\mathrm{NH_3}$\ inversion transitions are dominated by collisions rather than by radiative excitations and de-excitations, $T_\mathrm{rot}$ is a good measure for the kinetic temperature of the system (\citealt{malcolm:1983,danby:1988}). The background temperature was taken at 2.73\,K. With the total column density of $\mathrm{C^{18}O}$, $N_\mathrm{C^{18}O}$, known, we calculated the abundance of $\mathrm{C^{18}O}$\ relative to H$_2$, $\chi_\mathrm{C^{18}O}$, by comparing it to the hydrogen column density: \begin{equation} \chi_\mathrm{C^{18}O}=\frac{N_\mathrm{C^{18}O}}{N_\mathrm{H_2}}\,. \end{equation} The canonical number for $\chi_\mathrm{C^{18}O}$ in the nearby ISM is $1.7\times10^{-7}$ (\citealt{frerking:1982}). At larger distances towards the molecular ring and the Galactic center, this value is expected to be higher (\citealt{wilson:1994}), which would increase the CO depletion we measured. The ratio between the observed abundance, $\chi_\mathrm{obs}$, and the canonical value for the abundance, $\chi_\mathrm{canonical}$, then resulted in $\eta$ the CO depletion factor: \begin{equation} \eta=\frac{\chi_\mathrm{canonical}}{\chi_\mathrm{obs}}\,. \end{equation} All derived quantities, $N_\mathrm{C^{18}O}$, $N_\mathrm{H_2}$, $\chi_\mathrm{C^{18}O}$, and $\eta$ are listed in Table \ref{ta:c18o}. In most of the clumps, $\mathrm{C^{18}O}$\ appears to be depleted. We find a CO depletion of $\eta$=0.5--5.5 with an average of 2.3, which is slightly lower than what is found in clumps of IRDCs ($\eta$=2--10, \citealt{pillai:2007}). Observations by, e.g., \citet{kramer:1999} and \citet{bacmann:2003} have shown that molecular depletion is enhanced at lower temperatures. In Fig.~\ref{fig:eta-t}, we plot for all clumps the CO depletion against the rotational temperature. While \citet{kramer:1999} found an strong exponential dependence of CO depletion with temperature, our data are too scattered to differentiate between linear and exponential behavior. Nevertheless we can see that higher $\eta$ corresponds on average to colder sources (dashed black line). Figure \ref{fig:eta-t} shows that clumps in diffuse clouds (green crosses and green line) and peaked clouds (blue triangles and blue line) indeed follow the trend of higher CO depletion with colder temperatures. For clumps in multiply peaked clouds (red squares and red line), the situation is different; the clumps have a large spread in temperature as well as $\eta$, and on average the opposite trend is seen: increasing $\eta$ with higher temperature. This is mainly caused by a few clumps with high temperature and high $\eta$, G017.19+0.81 MM2, G014.63--00.63 MM1, and G022.06+0.21 MM1, all three of which have dust column densities above $10^{23}\,\mathrm{cm^{-2}}$ and water masers. Possibly these sources were heated so fast by the embedded YSO that the CO was not freed into the gas phase yet and the depletion values are thus still high. \begin{figure} \centering \includegraphics[angle=-90, width=6cm]{trot-eta-new-3.eps} \caption{\label{fig:eta-t} CO depletion versus the rotational temperature derived from $\mathrm{NH_3}$ . The green crosses represent sources in diffuse clouds, the blue triangles clumps in peaked clouds, and the red squares sources in multiply peaked clouds. The black dashed line marks the trend of an increase in CO depletion with colder temperatures for all the clumps, while the green, blue, and red solid lines mark the trends for the sources in diffuse, peaked, and multiply peaked clouds.} \end{figure} \section{Discussion} \subsection{Infall velocities} In Sect.~\ref{sect:infall}, we determined by binomial tests that the $\mathrm{HCO^+}$(1--0) line was the best suited of the molecular transitions that we had observed to study infall. For sources that had a double-peaked $\mathrm{HCO^+}$(1--0) line profile, indicating a systemic infalling motion, we calculated the infall velocity $v_\mathrm{in}$ according to \citet{myers:1996} as \begin{equation} v_\mathrm{in}= \frac{\Delta v_\mathrm{thin}^2}{v_\mathrm{red} - v_\mathrm{blue}}\ln\Big(\frac{1+e^{(T_\mathrm{blue}-T_\mathrm{dip})/T_\mathrm{dip}}}{1+e^{(T_\mathrm{red}-T_\mathrm{dip})/T_\mathrm{dip}}}\Big),\, \label{eq:vin} \end{equation} where $v_\mathrm{red}$ and $v_\mathrm{blue}$ are the LSR velocities of the red and blue peaks, whose main beam temperatures are given by $T_\mathrm{red}$ and $T_\mathrm{blue}$, respectively, and $T_\mathrm{dip}$ is the main beam temperature of the dip between the two peaks. All these quantities refer to the spectral shape of the optically thick line, except for $\Delta v_\mathrm{thin}$. With the volume density of the clump $n_\mathrm{H_2}$ and clump radius $R$, the infall velocity can be converted to a mass infall rate $\dot{M}_\mathrm{in}$ (after \citealt{myers:1996}): \begin{equation} \label{eq:min} \dot{M}_\mathrm{in}=4\pi R^2 n_\mathrm{H_2} \mu m_\mathrm{H}v_\mathrm{in}\,, \end{equation} where $\mu=2.33$ the mean molecular weight and $m_\mathrm{H}$ the mass of an hydrogen atom. For the clump $n_\mathrm{H_2}$ and clump radius $R$, we took the volume density and clump radius determined in Paper I from 1.2\,mm dust continuum observations. For low-mass star formation, the mass infall rates onto a protostar are between a few times $\sim10^{-5}\,M_\odot~\mathrm{yr^{-1}}$ (\citealt{palla:1993,kirk:2005}) and $\sim10^{-6}\,M_\odot~\mathrm{yr^{-1}}$ (\citealt{whitney:1997}). This is too low to form high-mass stars, since their prestellar phase is much shorter, less than $3\times10^4\,$yr, according to \citet{motte:2007}, which would mean a mass infall rate of $3\times10^{-4}\,M_\odot~\mathrm{yr^{-1}}$ to form a 10$,\,M_\odot$ star. Theoretical models predict a mass accretion onto a high-mass protostellar core (at radii $<0.1\,$pc) of $\dot{M}_\mathrm{in}$ of $10^{-3}\,M_\odot~\mathrm{yr^{-1}}$ (\citealt{mckee:2003,banerjee:2007}). Observations of massive clumps (with typical radii $>0.1$\,pc) find mass infall rates of $10^{-3}-10^{-4}\,M_\odot~\mathrm{yr^{-1}}$ (\citealt{fuller:2005,peretto:2006}). In high-resolution observations of high-mass protostellar cores, the mass infall rate is indeed higher and covers the predicted order of magnitude ($10^{-2}-10^{-4}\,M_\odot~\mathrm{yr^{-1}}$, \citealt{beltran:2006a,beltran:2008}). Our mass infall results, given in Table \ref{ta:infallv}, fall in the range of the observed values for massive star-forming clumps. To explore further the mass infall rate, we calculated the infall timescale using the clump mass obtained from the 1.2\,mm continuum (Paper I): $\tau_\mathrm{in}=\frac{M}{\dot{M}}$\,yr. This timescale can be compared with the free-fall timescale $\tau_\mathrm{ff} = 3.3\times10^7 \mathrm{n}^{-1/2}$\,yr, which was calculated using the density obtained from the same 1.2\,mm continuum observations (Paper I). Both timescales are given in Table~\ref{ta:infallv}. In general, the free-fall timescale is longer than the estimated infall timescale, which indicates that the mass infall rate calculated in Eq.~\ref{eq:min} is likely overestimated. The formula in Eq.~\ref{eq:min} assumes a spherically symmetric infall of the entire clump, while most likely the $\mathrm{HCO^+}$(1--0) emission is originating in an optically thick low-density layer in the outer parts of the clump (see also \citet{sepulcre:2010}). We also calculated the $\mathrm{HCO^+}$(4-3) infall velocities and mass infall rates for the few sources that showed a double-peaked blue-excess line profile (Table~\ref{ta:infallv}). Unfortunately, the error in the infall velocity of G024.94--00.15\,MM1 was quite high, leaving only G014.63--00.57\,MM1 to serve as a comparison. For the latter, the infall velocity derived from the higher excitation $\mathrm{HCO^+}$(4--3) line profile is half of the one derived from the $\mathrm{HCO^+}$(1--0) line profile, and the $\mathrm{HCO^+}$(4--3) infall timescale is twice the $\mathrm{HCO^+}$(1--0) timescale. At least for this one source, we find an indication that the $\mathrm{HCO^+}$(1--0) infall velocities and mass infall rates are indeed overestimated. \begin{table*} \centering \caption{Infall velocities, mass infall rates, and timescales for the blue excess sources \label{ta:infallv}} \begin{tabular}{llccccccr} \hline\hline \noalign{\smallskip} Source & &\multicolumn{1}{c}{$v_{\mathrm{in10}}$}& \multicolumn{1}{c}{$\dot{M}_{\mathrm{in10}}$} & \multicolumn{1}{c}{$\tau_\mathrm{in10}$} & \multicolumn{1}{c}{$v_{\mathrm{in43}}$}& \multicolumn{1}{c}{$\dot{M}_{\mathrm{in43}}$} & \multicolumn{1}{c}{$\tau_\mathrm{in43}$} &$\tau_\mathrm{ff}$\\ & &\multicolumn{1}{c}{($\mathrm{km~s^{-1}}$)} & \multicolumn{1}{c}{($10^{-3} M_\odot~\mathrm{yr^{-1}}$)} & \multicolumn{1}{c}{($10^5$\,yr)} & \multicolumn{1}{c}{($\mathrm{km~s^{-1}}$)} & \multicolumn{1}{c}{($10^{-3} M_\odot~\mathrm{yr^{-1}}$)} & \multicolumn{1}{c}{($10^5$\,yr)} &($10^5$\,yr)\\ \noalign{\smallskip} \hline \multicolumn{6}{l}{{\em Diffuse clouds}}\\ G013.28--00.34... &MM1& 1.09(0.15) & 1.5 & 0.6 & - & - & -&1.5 \\ G013.97--00.44... &MM1& 0.31(0.13) & - & - & -& - & - & -\\ \multicolumn{6}{l}{{\em Peaked clouds}}\\ G013.91--00.51... &MM1& 0.69(0.10) & 2.2& 0.66 & - & - & -&0.70\\ G024.94--00.15... &MM1& 1.34(0.12) &3.3 &0.32 &0.34(0.66) & 0.85 & 1.2&0.74 \\ &MM2& 3.00(0.13) &6.2 &0.13& - & - & -& 0.78\\ G030.90+00.00... &MM2& 1.76(0.15) &9.4&0.75& - & - & -&1.50\\ G035.49--00.30... &MM1& 7.4(0.13) &16.6 &0.06 & - & - & -& 0.78\\ &MM2& 0.84(0.10) &1.4 &0.5& - & - & -&0.85\\ G053.81--00.00... &MM1& 7.4(0.10) &9.6 &0.03& - & - & -&0.55 \\ \multicolumn{6}{l}{{\em Multiply peaked clouds}}\\ G014.63-00.57... & MM1 & 2.7(0.14) & 22.0 & 0.15& 1.30(0.35) & 10.6 & 0.30& 0.38 \\ & MM2 & - & - & - &0.90(0.28) & 2.7 & 0.26 & 0.38\\ G017.19+00.81... &MM2& 0.66(0.20) &3.1 & 0.46& - & - & -&0.42\\ &MM3& 1.65(0.13) &2.8& 0.26& - & - & -&0.85\\ G018.26--00.24... &MM3& 6.1(0.18) &16.0& 0.08& - & - & -&0.80\\ \noalign{\smallskip} \hline \end{tabular} \tablefoot{Columns are (from left to right) source name, infall velocity; mass accretion rate; accretion timescale based on the $\mathrm{HCO^+}$(1--0) line, infall velocity; mass accretion rate; accretion timescale based on the $\mathrm{HCO^+}$(4--3) line, and free-fall timescale.} \end{table*} \subsection{Evolutionary sequence} \label{sec:evodis} In Paper I, we defined three classes of high extinction clouds, based on the contrast of the clump to the cloud in the 1.2\,mm continuum and the number of clumps per cloud. These clouds comprise diffuse clouds, clouds with a low clump-to-cloud column density contrast $C_{N_\mathrm{H_2}}$$<$2, peaked clouds, clouds with higher contrast 2$<$$C_{N_\mathrm{H_2}}$$<$3 having at least one clump, multiply peaked clouds, clouds with still higher contrast $C_{N_\mathrm{H_2}}$$>$3 that generally contained several clumps. Here, we review all the results in light of the hypothetical evolutionary sequence we proposed, i.e., from the quiescent diffuse clouds to the peaked ones, and finally to the multiply peaked clouds. \subsubsection{Diffuse clouds} For the class of diffuse clouds, more than half of the investigated sources were not well-defined clumps but positions of diffuse emission within a cloud. The skewness parameter analysis found a slight excess of blue sources, indicating infall, but the binomial test showed that this result is not significant (see Table \ref{ta:bin}). The diffuse clouds are likely the phase where the clumps are beginning to collapse. This would agree with the clump masses found in Paper I, where on average we found less massive clumps in the diffuse clouds than in the peaked or multiply peaked clouds (see also Table\,\ref{ta:sample}). Moreover, few YSO/hot core indications were found toward clumps in diffuse clouds, and half of the starless and prestellar sources were found toward the diffuse clouds (the other half was found toward the peaked clouds). A young evolutionary stage can also be identified by the CO depletion, since CO depletion requires a certain density ($>$ few$10^4\,\mathrm{cm^{-3}}$, \citealt{kramer:1999,bergin:2007}). Low depletion levels can indicate that the clump has not yet reached a high enough density. The mean depletion level of clumps in diffuse clouds ($\eta$=1.9) was slightly lower than that found toward the clumps in peaked and multiply peaked clouds ($\eta$=2.3). The depletion timescale, $\tau_\mathrm{d}\sim5\times10^9/n_\mathrm{{H_2}} ($\citealt{bergin:2007}), is $\sim$$4\times10^4$\,yr for clumps in diffuse clouds, which is slightly longer than that for in peaked clouds ($\sim$$3\times10^4$\,yr) and multiply peaked clouds ($\sim$$2\times10^4$\,yr). \begin{table}[!h] \centering \caption{\label{ta:bin} Infall properties from the $\mathrm{HCO^+}$(1--0) line profile for different classes of clouds} \begin{tabular}{lrrrrrrr} \hline\hline \noalign{\smallskip} Class & $N_\mathrm{blue}$ & $N_\mathrm{red}$ & $N_\mathrm{skew}$ & $N_\mathrm{tot}$ & $E$ &$P$ & $<\delta v>$\\ \noalign{\smallskip} \hline \noalign{\smallskip} Diffuse & 4 & 2 &6 & 8 & 0.25 & 0.34 & $-0.001$\\ Peaked & 11 & 2 &13 & 17 & 0.53 & 0.01 & $-0.63$\\ Multiply p. & 9 & 8 &17 & 20 & 0.05 & 0.50 & $-0.72$\\ \noalign{\smallskip} \hline \end{tabular} \tablefoot{Columns are (from left to right) number of blue excess sources, number of red excess sources, total number of sources with significant $\mathrm{HCO^+}$(1--0) skewness, total number of sources, the excess parameter, the probability of the distribution to arise by chance, and the average $\mathrm{HCO^+}$(1--0) skewness.} \end{table} \subsubsection{Peaked clouds} In the peaked clouds, the distribution of blue and red sources from the $\mathrm{HCO^+}$(1--0) line profile significantly favors the blue sources with infall. The excess parameter was 0.53 with only a 1\% possibility for this distribution to arise by chance (Table \ref{ta:bin}). Apparently, most clumps in the peaked clouds are accreting material. The infall velocities and mass accretion rates for peaked (as well as the one for multiply peaked) clouds have a very wide range, whose mean lies far above the infall velocity/mass accretion rate of the diffuse clouds. All the SiO detections of clumps in peaked clouds coincide with sources that show infall based on the $\mathrm{HCO^+}$(1--0) skewness parameter (Fig.~\ref{fig:sio-infall}). The strong correlation between the infall characteristics and SiO outflows shows that accretion is accompanied by collimated outflows, which leads to an enhancement of SiO. This correlation does not exist for the earlier or later classes. \subsubsection{Multiply peaked clouds} Among the multiply peaked clouds, there were almost equal numbers of blue and red excess sources (based on the $\mathrm{HCO^+}$(1--0) skewness). These clouds seem to be hosting clumps in a variety of evolutionary stages: from clumps with still ongoing accretion to clumps where the feedback from the formed young stars halts further accretion. Many clumps in these clouds presented SiO emission. Figure \ref{fig:sio-infall} shows that the number of clumps with SiO emission showing infall is about equal to those showing expansion (both infall and expansion characteristics are based on the $\mathrm{HCO^+}$(1--0) skewness parameter), meaning that the outflow continues even after large-scale infall has finished. The volume density of the multiply peaked clouds, $n_\mathrm{H_2}>10^5\,\mathrm{cm}^{-3}$, is much larger than what is found for the peaked clouds. Multiply peaked clouds have the highest detection rate of the $\mathrm{CH_3CN}$(5--4) $K$=0 and $\mathrm{H_2CO}$(4$_{03}$--3$_{04}$) lines, indicating that this class contains the most YSOs or hot molecular cores. In Paper I, we already found that the $\mathrm{NH_3}$\ rotational temperatures were on average higher for clumps in multiply peaked clouds than for the other classes (Table\,\ref{ta:sample}). The $K$ levels of the $\mathrm{CH_3CN}$(5--4) line allowed a handful of rotational temperature estimates, ranging between 23 and 34\,K. The increasing evidence for YSOs is likely the reason for the broadening of the line width found toward clumps in this class (observed for the $\mathrm{N_2H^+}$(1--0), $\mathrm{NH_3}$(1,1), $\mathrm{C^{18}O}$(2--1), and $\mathrm{H^{13}CO^+}$(1--0) lines) by increasing turbulence. Finally, three clumps in the multiply peaked clouds, G014.63-00.57 MM1, G017.19+00.81 MM2, and G022.06+00.21 MM1, all show various characteristics that are different to the properties of the bulk of the sample. Examples include their SiO intensity against column density, the CO depletion against temperature, and their $\mathrm{N_2H^+}$(1--0)/$\mathrm{N_2H^+}$(3--2) ratio. While these clumps have strong outflows, suggested by their SiO emission and the presence of water masers, they also have very high column densities (theoretically enough to form massive stars, e.g., \citealt{krumholz:2008}). They show strong CO depletion values while having elevated temperatures ($\sim$ 20\,K) due to the rapid heating of the environment by the YSO. Most likely, these three clumps are actively forming massive YSOs, which is why they have such extreme properties. \section{Summary} We performed spectral line observations of several molecules using the IRAM 30m and the APEX telescopes toward a sample of clumps in high extinction clouds spanning various evolutionary stages. The skewness parameter was used to classify sources based on their line profiles as infall sources or outflow/expanding sources. The $\mathrm{HCO^+}$(1--0) and CO(3--2) lines were found to be more sensitive to infall and outflow than the $\mathrm{HCO^+}$(4--3) line. Only the $\mathrm{HCO^+}$(1--0) profiles showed a significant excess of infall sources, and therefore the $\mathrm{HCO^+}$(1--0) line profile was found to be most effective to investigate infall profiles. SiO emission, which is thought to be a tracer of shocks, was predominantly detected toward sources with infall. This finding is consistent with a scenario in which accretion is accompanied by collimated outflows that lead to an enhancement of SiO. Between the three cloud classes, the peaked clouds have a significant high excess of infall sources, while for the other classes no significant excess was found. Possibly, the infall phase has not yet begun in most diffuse clouds and already has come to an end in some of the clumps of the multiply peaked clouds. The infall velocities and mass infall rates suggest an increasing trend when going from the diffuse clouds (1.5 $\times10^{-3}\,M_\odot~\mathrm{yr^{-1}}$) to the peaked (and multiply peaked) clouds (1.5--16$\times10^{-3}\,M_\odot~\mathrm{yr^{-1}}$) suggesting that mass infall increases in these later stages. More measurements of clumps in diffuse clouds are, however, necessary to confirm this. From RADEX calculations, we found that the ratio of the $\mathrm{N_2H^+}$(1--0) to $\mathrm{N_2H^+}$(3--2) integrated intensity is sensitive to the hydrogen volume density, and together with the $\mathrm{N_2H^+}$(1--0) optical depth can be used to derive good estimates of not only the hydrogen volume density but also the intrinsic $\mathrm{N_2H^+}$\ column density. Apart from the ammonia rotational temperatures, determined in Paper I, we added several other temperature estimates. All estimates converge to a temperature range between 10 and 40\,K, where the higher value is determined by $\mathrm{CH_3CN}$, a molecule usually found in hot cores and UCH{\sc ii}\,regions. Similarly, various density estimates place the density in the order of $10^5\,\mathrm{cm^{-3}}$. The presence of YSOs in the clumps of the high extinction clouds was established by 24\,$\mu$m emission, but also indirect evidence through outflows (SiO emission) and the presence of hot molecular cores through $\mathrm{CH_3CN}$\ and $\mathrm{H_2CO}$\ emission was gathered. We find that most of the YSOs are located in clumps of peaked and multiply peaked clouds, while almost no YSO indications were present in diffuse clouds. The presence of an YSO seems independent of the physical size of the clump, since YSOs are found over a large range of clump diameters. However, there seems to be (hydrogen) column density cutoff at $(4-5)\times10^{22}\,\mathrm{cm^{-2}}$, below which few or no YSOs are found. The cutoff is most clear in the 24\,$\mu$m emission and in the presence of SiO emission. In summary, the diffuse clouds seem to contain clumps that are on the verge of infall. The first stages of star formation are ongoing in the peaked clouds, since a majority of the clumps is infalling and many clumps also show outflows. The multiply peaked clouds harbor the more evolved clumps, where infall has already stopped in some of them, and generally we find indications of YSOs. Hence, the cloud morphology seems to have a direct connection with the evolutionary stage of the objects forming inside. \begin{acknowledgements} We would like to express our thanks to Floris van der Tak and Malcolm Walmsley for their constructive comments that lead to a much improved manuscript. We are grateful for the help and assistance of the staff of both the IRAM 30m and the APEX 12m telescopes during the observations. KLJR is funded by an ASI fellowship under contract number I/005/11/0. \end{acknowledgements} \bibliographystyle{aa}
\section{Introduction} There have been a variety of approaches to the problem of characterizing what is non-classical about quantum theory. There are many important signatures of quantum theory, but with the rise of quantum information, the exponential speedups in quantum algorithms over the best known classical algorithms have increasingly become an important signature of quantum theory. This point is especially poignant in light of recent work by Aaronson and Arkhipov wherein a simple non--universal linear optical system is shown to be able to perform computational tasks believed to be hard for classical computers~\cite{Aaronson2010Computational}. The extent to which computational speedups and the boundaries between computational complexity classes are reflected in more traditional measures of non--classicality remains, however, an open question. In continuous variable quantum theory, and quantum optics in particular, the most frequently considered notions of quantumness are phrased in terms of the so-called quasi-probability distributions, such as the Wigner function and the (Glauber-Sudarshan) $P$-function. There is a strong tradition in physics of considering negativity of the quasi-probability function as an indicator of non--classicality of a quantum state \cite{Mandel1986NonClassical,Paz1993Reduction,Bell2004Speakable,Kalev2009Inadequacy}. It is therefore natural to suspect that negativity is intimately linked to computational speedups in both discrete and continuous quantum information processing. Continuous variable quantum information theory provides a potentially powerful alternative to the usual discrete formalism and many of the seminal results in discrete variable quantum computation have analogs in the continuous variable setting. Perhaps the most important example is the {}``continuous variable Gottesman-Knill theorem'', which states that a computation restricted to the subset of quantum theory containing only Gaussian states and operations is classically efficiently simulatable \cite{Bartlett2002Efficient,Bartlett2002Efficient2}. More concretely, unitary Gaussian quantum information is defined to be the following set of operators (see, for example, \cite{Weedbrook2011Gaussian}): $n$ mode Gaussian input state; quadratic Hamiltonians; and, measurements with (or without) post-selection onto Gaussian states. Bartlett \emph{et al.} \cite{Bartlett2002Efficient,Bartlett2002Efficient2} have shown explicitly that there exists a classical algorithm that reproduces the output probabilities of the measurement results that executes in time that scales polynomially with the number of modes. This shows that some non-Gaussian resources are necessary to obtain an exponential speed-up with quantum optical experiments, but leaves open the question of whether they are sufficient. Recently, Veitch \emph{et al.}~\cite{Veitch2012Negative} have shown that a discrete analog of the Wigner function ~\cite{Gross2006Hudsons} can be used to define a necessary condition for a mixed quantum state to enable an exponential speed-up through quantum computation. Their model considers the use of Clifford operations on \emph{qudits} (the case of qubits is not covered by their proof) and measurements of stabilizer states and finds, somewhat surprisingly, that there exist a class of \emph{bound universal states} outside of the convex set of stabilizer states that can still be efficiently simulated and therefore do not serve as a resource for exponential speed-up with quantum computation. There is a tight mathematical correspondence between the discrete and continuous Wigner representations, the Clifford/stabilizer model for qudits \cite{Gottesman1997Stabilizer} and the Gaussian model for quantum optics considered by Bartlett \emph{et al.}~\cite{Bartlett2002Efficient,Bartlett2002Efficient2}. It is therefore natural to ask whether the restriction to Gaussian states in the model of Bartlett \emph{et al.} can be relaxed to allow more general class of initial states that have non-negative Wigner representation while still permitting an efficient classical simulation. This work affirms an answer in the positive by showing that a large class of quantum states with positive Wigner representation exists outside the convex hull of the $n$-mode Gaussian states that can be efficiently simulated using a classical computer, given restrictions to quadratic Hamiltonians and Gaussian measurement. This shows, in particular, that linear optical quantum devices are essentially no more computationally powerful than classical computers under such restrictions. We show this by providing an explicit classical simulation algorithm that can be used to simulate sampling the output probability distributions of the evolved initial states. As a practical application we apply our results to determine a threshold on the computational power of single-photon-added-thermal states (SPATS) \cite{Agarwal1992Nonclassical,Zavatta2004QuantumtoClassical,Parigi2007Probing} for variable efficiencies. In this sense, our work serves as both a conceptual and practical generalization of the continuous variable Gottesman--Knill theorem to a broader class of input states. This paper is outlined as follows. We begin with a brief review of the Wigner function formalism and Gaussian quantum mechanics in section~\ref{sec:wigner}. We then provide our simulation protocol for states with positive Wigner representation in section~\ref{sec:sim}. In section~\ref{sec:SPATS}, we discuss positivity of the Wigner function as a necessary condition for quantum computation. We illustrate the bound state region via the recently studied class of limited-efficiency SPATS and show that quantum efficiencies of 50\% are a necessary threshold for computational speed-up. Finally, section~\ref{sec:conclusion} contains our conclusion and further discussion about our findings. \section{Review of Wigner Functions}\label{sec:wigner} Wigner functions provide a natural quantum analog of the classical phase space distribution of a dynamical system. We provide below a brief review of the properties of Wigner functions. For simplicity, we focus our attention on Wigner functions for a single particle (or equivalently a single mode) in one dimension. The generalization to higher dimensions and more particles is straightforward~\cite{Wigner1932On}. The Wigner representation of a state $\rho$ is defined to be~\cite{Wigner1932On} \begin{equation} W_{\rho}(q,p)=(2\pi)^{-1}\int_{-\infty}^{\infty}\bra{q-y/2}\rho\ket{q+y/2}e^{ipy/\hbar}\mathrm{d}y, \end{equation} where $\ket q$ is a position eigenstate. The Wigner function is both positive and negative in general. However, it otherwise has many of the same properties as a classical probability density on phase space. For these reasons, the Wigner function is often referred to as a \emph{quasi-probability} function. Intuitively, if we could find a bona fide joint probability distribution of non-commuting observables, then there would be no difference between quantum and classical theories. It is not surprising, then, that \emph{negativity} is necessary in all possible quasi-probability representations of a quantum state \cite{Ferrie2010Necessity}. The time--evolution of the Wigner function for a Hamiltonian of the form $H=p^{2}/2m+V(q)$ is given by~\cite{Wigner1932On,balbook} \begin{equation} \frac{\partial W_{\rho}(q,p)}{\partial t} =\{H,W_\rho\} + \sum_{\ell=1}^\infty\frac{1}{(2\ell+1)!}\left(-\frac i2\right)^{2\ell}\frac{\partial^{2\ell+1}V(q)}{\partial q^{2\ell+1}}\frac{\partial^{2\ell+1}W_{\rho}(q,p)}{\partial {p}^{2\ell+1}}, \end{equation} where $\{\cdot,\cdot\}$ is the Poisson bracket, which governs classical Hamiltonian equations of motions. This result is important because it states that the time--evolution of $W_{\rho}(q,p)$ is given by Liouville's equation, plus a quantum correction. The quantum correction is zero for the case of the quadratic Hamiltonian: \begin{equation}\label{eq:Wpevol} \frac{\partial W_{\rho}(q,p)}{\partial t} = \{H,W_\rho\}. \end{equation} Hence the evolution equation agrees \emph{precisely} with the classical predictions. This observation will be vital for our simulation algorithm because the Hamiltonians permitted by linear optics are quadratic (Harmonic oscillators), which (along with our non--negativity assumption) implies that we will be able to simulate the evolution of $W_{\rho}(q,p)$ using an ensemble of classical trajectories. This discussion above implies that a Wigner function that is initially classical, meaning that it is non--negative and hence interpretable as a probability density function (pdf), will remain classical under the action of a quadratic Hamiltonian. In this context is therefore useful to determine the conditions under which a Wigner function is non--negative as this gives a practically relevant boundary between quantum and classical states. Hudson's theorem \cite{Hudson1974When} was the first attempt to characterize the positive Wigner functions and it was later generalized to the following \cite{SotoEguibar1983When}. Let $\psi$ be a pure quantum state of $n$ oscillators (modes). Then its Wigner function is positive if and only if \begin{equation} \psi(\vec{Q})=e^{-\frac{1}{2}(\vec{Q}\cdot A\vec{Q}+B\cdot\vec{Q}+c)}, \end{equation} where $A$ is an $n\times n$ Hermitian matrix, $B$ is an $n$-dimensional complex vector and $c$ is a normalization constant. In quantum optics terminology, these are either \emph{coherent states} or \emph{squeezed states}. That is, plugging these states into the definition of the Wigner function yields multivariate Gaussian distributions in phase space. Convex combinations of these states (incoherent mixtures of them) also have positive Wigner function since the mapping is linear. Early on, these were incorrectly conjectured to be the only such mixed states with positive Wigner function. The question of mixed states was given a full treatment in reference \cite{Srinivas1975Some} and later in \cite{Brocker1995Mixed}. Both references independently found that a theorem in classical probability attributed to Bochner \cite{Bochner1933Monotone} and generalization thereof can be used to characterize both the valid Wigner functions and the subset of positive ones. What was shown is that there exist a large class of states with positive Wigner function that are not convex combinations of Gaussian states. So far, these states have received little attention. In Section~\ref{sec:SPATS}, we show that such states are more than a mathematical curiosity; they arise naturally in quantum optics. Gaussian measurements are also easily modeled in the Wigner representation. Recall that for non-negative states the Wigner function picture allows us to represent the system as a probability density over underlying physical states in phase space, $\bm{u}_{f}$. Gaussian measurements in this picture are also modeled as probability densities for outcomes $\bm{k}$, conditioned on the value of the underlying physical state $\bm{u}_{f}$. Specifically, consider the case of measurement $\mathcal{M}$ of a Gaussian state $G$ with covariance matrix $V_\mathcal{M}$. The POVM representation of this measurement is \cite{Leonhardt1998Measuring} \begin{equation} \mathcal{M}(V_{\mathcal{M}})=\{\ketbra{G(\bm{k},V_{\mathcal{M}}}{G(\bm{k},V_{\mathcal{M}})}\ :\ \bm{k}\in\mathbb{R}^{2n}\},\label{eq:guass_meas_povm} \end{equation} which selects a Gaussian state with mean $\bm{k}$ and covariance $V_\mathcal{M}$ from all possible Gaussian states with mean $k$ and covariance $V_\mathcal{M}$. In the Wigner function picture the representation of this measurement is, \begin{equation} M_{\mathcal{M}(V_{\mathcal{M}})}(\bm{k}|\bm{u}_f)=\frac{1}{\mathcal{X}}\exp\left(-(\bm{k}-\bm{u}_f)^{T}V_\mathcal{M}^{-1}(\bm{k}-\bm{u}_f)\right),\label{eq:gauss_meas_wig_rep} \end{equation} where $\mathcal{X}$ is the normalization constant. We introduce the notation in~\eqref{eq:gauss_meas_wig_rep} to emphasize the difference between the representations of measurements and states. The interpretation of this equation is that if the system is actually at the point $\bm{u}_f$ the effect of a measurement will be to produce an outcome $\bm{k}$ according to the probability density $M_{\mathcal{M}(V_{\mathcal{M}})}(\bm{k}|\bm{u}_f)$. Using this equation and the law of total probability we can find the probability density of measurement outcomes $\bm{k}$ for the measurement $\mathcal{M}(V_{\mathcal{M}})$ on a quantum state with Wigner representation $W_{\rho}(\bm{k})$: \begin{equation} p(\bm{k}|\mathcal{M}(V_{\mathcal{M}}),\rho)=\int_{\bm{u}}W_\rho(\bm{u})M_{\mathcal{M}(V_{\mathcal{M}})}(\bm{k}|\bm{u})d\bm{u}. \end{equation} Of course, this agrees with the probability assigned by the Born rule \begin{equation} p(\bm{k}|\mathcal{M}(V_{\mathcal{M}}),\rho)=\text{Tr}(\ketbra{G(\bm{k},V_{\mathcal{M}}}{G(\bm{k},V_{\mathcal{M}})}\rho). \end{equation} The simulation algorithm that we propose uses none of the special properties of Gaussian measurements other than the fact that they have a positive Wigner representation and that we can efficiently draw samples from a Gaussian distribution. This means that our results will apply to any measurements that satisfy these properties. We focus our attention on Gaussian measurements rather than these more general measurements because of their simplicity and physical relevance. \section{Simulation Algorithm}\label{sec:sim} At first glance it may seem that a simulation algorithm for linear optics may be difficult owing solely to exponential size of the Hilbert space dimension that is generated by the evolution. We overcome this problem by exploiting the fact that our Hamiltonians are quadratic in $p$ and $q$, which implies that the evolution of the Wigner function follows the Liouville equation as shown in Eq.~\eqref{eq:Wpevol}. Since the Liouville equation preserves non-negativity and probability mass, the Wigner function will remain a classical distribution throughout the evolution. This allows us to model evolution of the Wigner function using an ensemble of classical trajectories, each of which can be efficiently simulated. The resulting trajectories can be used to efficiently draw samples from the final distribution of measurement outcomes prescribed by the Born rule without needing to know the final quantum state. It is critical to understand that we are not simulating the evolution of the quantum state, rather we are simulating measurement outcomes from a quantum circuit; this is exactly analogous to the difference between knowing a probability distribution and being able to sample from it. In particular, the ability to efficiently draw samples does not imply the ability to efficiently learn the underlying distribution because the dimension of the probability distribution on $n$ modes is exponentially large. We show in this section that this simulation strategy can be used to efficiently sample from the output of the following class of quantum algorithms: \begin{algorithmclass}[H] \begin{enumerate} \item Apply the linear optical transformation $U_{T,\bm{x}}$. \item Perform the separable Gaussian measurement $\mathcal{M}(V_{\mathcal{M}})=\measj{1}\otimes\measj{2}\otimes\dots\otimes\measj{n}$, where we follow the naming convention of equation \ref{eq:guass_meas_povm} and the tensor product is understood to mean that the POVM elements of $\mathcal{M}(V_{\mathcal{M}})=\measj{1}\otimes\cdots\otimes\measj{n}$ are tensor product combinations of the POVM elements of $\mathcal{M}(V_{\mathcal{M}})$ in the obvious way. \item Return the measurement outcome $\bm{k}=(\bm{k}_{1},\bm{k}_{2},\cdots,\bm{k}_{n})\in\mathbb{R}^{2n}$ corresponding to the mean of a Gaussian POVM element. \end{enumerate} \caption{Family of efficiently simulatable quantum algorithms\\ \textbf{Input:} Number of modes $n$, an initial $n$ mode quantum state $\rho=\rho_1\otimes\dots\otimes\rho_{n}$ where each $\rho_j$ has positive Wigner representation $W_{\rho_j}(\bm{u})$, a description of a linear optical transformation $U_{T,x}$ which is parameterized by $T\in \mathbb{R}^{2n\times 2n}$ and $\bm{x}\in\mathbb{R}^{2n}$.\\ \textbf{Output:} A string of measurement outcomes $\bm{k}$ sampled according to the probability density $p(\bm{K}_{\text{quant}}=\bm{k})$ determined by the Born rule. \label{alg:Simulable-Quantum-Circuit}} \end{algorithmclass} {\flushleft Here we conceive of any quantum algorithm in this class as a way of sampling outcome strings $\bm{k}$ distributed according to the probability densities given by the Born rule. We label the corresponding random variable $\bm{K}_{\text{quant}}$. Here we are \emph{not} simulating the evolution of the Wigner distribution, which would be equivalent to simulating the full quantum state. Rather, we show that there is a corresponding classical algorithm that produces outcome strings $\bm{k}$ with (very nearly) the same distribution those from algorithm class~\ref{alg:Simulable-Quantum-Circuit}.} Using intuition similar to that in~\cite{Veitch2012Negative}, we note that quantum algorithms in class~\ref{alg:Simulable-Quantum-Circuit}, can be simulated using the following classical algorithm, provided access to classical resources with infinite numerical precision and a blackbox function that can be used to draw samples from $W_{\rho_j}(\bm{u})$ for $j=1,\ldots,n$. We will later provide an algorithm that does not require infinite precision, but we provide the infinite precision algorithm first because it conveys the necessary intuition without focusing on the technical issues that arise when discretizing the distributions. \begin{algorithm}[H] \begin{enumerate} \item Sample $\bm{u}\in\mathbb{R}^{2n}$ according to the distribution $W_{\rho}(\bm{u})=W_{\rho_1}(\bm{u}_{1})\cdots W_{\rho_n}(\bm{u}_{n})$ by drawing a sample independently from each mode using the blackbox function. \item Apply the affine transformation $\tilde{\bm{u}}=T\bm{u}+\bm{x}$ corresponding to the linear optical transformation $U_{T,\bm{x}}$ to the sampled phase space point $\bm{u}$. This transformation is an affine mapping due to Louiville's theorem. \item Return the outcome string $\bm{k}=(\bm{k}_{1},\bm{k}_{2},\cdots,\bm{k}_{n})\in\mathbb{R}^{2n}$ from the distribution \\$M_{\mathcal{M}(V_{\mathcal{M}})}(\bm{k}|\tilde{\bm{u}})=M_{\measj{1}}(\bm{k}_{1}|\tilde{\bm{u}}_{1})\cdots M_{\measj{n}}(\bm{k}_{n}|\tilde{\bm{u}}_{n})$, where $M_{\measj{j}}(\bm{k}_{i}|\tilde{\bm{u}}_{i})$ is given as in equation \ref{eq:gauss_meas_wig_rep}, \[ M_{\measj{j}}(\bm{k}_{i}|\tilde{\bm{u}}_{i})=\frac{1}{\mathcal{X}_j}\exp\left(-(\bm{k}_{i}-\tilde{\bm{u}}_{i})^{T}V_{\mathcal{M},j}^{-1}(\bm{k}_{i}-\tilde{\bm{u}}_{i})\right). \] \end{enumerate} \caption{Infinite precision classical simulation algorithm for algorithms in class~\ref{alg:Simulable-Quantum-Circuit}\label{alg:Equivalent-Classical-Circuit}\protect \protect \\ \textbf{Input:} As algorithms in class~\ref{alg:Simulable-Quantum-Circuit}, except $\rho$ is not provided.\\% but is replaced with samples from $W_{\rho_j}(p,q)$.\\ \textbf{Output:} A string of measurement outcomes $\bm{k}$ sampled according to the probability density $p(\bm{K}_{\text{class}}^{(2)}=\bm{k})$.} \end{algorithm} {\flushleft The intuition behind this class of algorithms is to use the classical phase space model afforded to us by the non-negative Wigner functions and quadratic evolutions in order to turn the quantum problem into one that can be efficiently simulated by a classical computer. In this context we can think of our quantum system as actually being definitely at some point $\bm{u}\in\mathbb{R}^{2n}$ which is unknown to us. The point then moves under a fully deterministic classical evolution and measurement on each register amounts to picking a point $\bm{k}$ from a normal distribution centered at the location of the system. Each classical algorithm samples outcomes $\bm{k}$ according the probability density \[ p(\bm{K}_{\text{class}}^{(2)}=\bm{k})=\int_{\bm{u}}W_\rho(\bm{u})M_{\mathcal{M}(V_{\mathcal{M}})}(\bm{k}|\bm{u})d\bm{u}, \] which agrees with the density given by the Born rule. Thus the outcomes $\bm{K}_{\text{class}}^{(2)}$ from the classical simulator are distributed in exactly the same way as $\bm{K}_{\text{quant}}$, which are outcomes drawn from the actual quantum system.} Unfortunately, algorithms similar to~\ref{alg:Equivalent-Classical-Circuit} cannot be executed precisely on digital computers and instead would require an analog computer (often referred to as a \emph{real computer}). If physical, such computers would have unrealistic computational powers such as being able to solve NP--complete problems in polynomial time~\cite{VSD86} and would also violate the holographic principle~\cite{Aar05}. For these reasons, we need to discretize~\ref{alg:Equivalent-Classical-Circuit} in order to assess the cost of simulating linear optics on realistic classical computers. The major technical difference between the continuous variable case and the discrete case considered in \cite{Veitch2012Negative} involves showing that finite-precision errors can be made negligible with efficient overhead costs given a set of reasonable assumptions about the input states, dynamics and measurements. Since infinite precision is requried in the continuous variable setting, in order to specify a quantum state we must make some finite precision truncation. To this end we shall assume access to a family of oracles $\mathcal{W}_{\rho_{j},\eta}(l,m)$ that takes integers $l,m$ and returns a value satisfying $|\mathcal{W}_{\rho_{j},\eta}(l,m)-W_{\rho_{j}}(\bm{\mu}_{\rho_j}+(l,m)\delta)|<\eta.$ That is, each oracle queries the Wigner function at points on a grid centered at the mean of the distribution. This is a weak assumption as it does not require us to even know the state we are simulating. Using this resource the algorithm can be written as: \begin{algorithm}[H] \caption{Finite-precision classical simulation algorithm for quantum algorithms in class~\ref{alg:Simulable-Quantum-Circuit}\label{alg:advanced-Equivalent-Classical-Circuit}\\ \textbf{Input:} As algorithm~\ref{alg:Equivalent-Classical-Circuit}, but also require $\delta$, a discretization length for the input, $|\bm{\epsilon}_2|$, a bound for the numerical error involved in applying the affine transformation, $\Gamma$, a discretization length for the output, $|\mathcal{A}|$, the area truncated square region of phase space that the simulator considers and $\bm{\mu}_{\rho_j}$, the mean of the Wigner distribution of the quantum state on each mode $j$. We require $\sqrt{|\mathcal{A}|}$ to be an odd integer multiple of $\delta$ and $\Gamma$ to be an odd integer multiple of $\delta$.\\ \textbf{Output:} A string of measurement outcomes $\bm{k}$ sampled according to $\text{Pr} (\bm{K}_{\text{class}}=\bm{k})\equiv\text{Pr} _{\text{sim}}(\bm{k})$.} \begin{enumerate} \item For each $j=1,\ldots, n$ execute a through d. (This step approximates sampling a point from phase space.) \begin{enumerate} \item For each integer $l,m \in \left[-\frac{\sqrt{|\mathcal{A}|}}{\delta},\frac{\sqrt{|\mathcal{A}|}}{\delta}\right]$ set $\text{Pr} _{\text{{\rm sim}},\rho_{j}}(l,m)=\mathcal{W}_{\rho_{j},\eta}(l,m)\cdot\delta^{2}.$ \\(The phase space is truncated to a size $|\mathcal{A}|$ and discretized into boxes of size $\delta$. This step sets a pdf over the centers of the boxes.) \item For each $(l,m)$ set $\text{Pr} _{\text{{\rm sim}},\rho_{j}}(l,m)=\text{Pr} _{\text{{\rm sim}},\rho_{j}}(l,m)/\sum_{l,m}\text{Pr} _{\text{sim},\rho_{j}}(l,m)$ (This step normalizes the pdf.) \item Draw a sample $(l,m)$ from the pdf $\text{Pr} _{\text{{\rm sim}},\rho_{j}}(l,m).$ \item Set $\boldsymbol{u}_{j}=\boldsymbol{\mu}_{j}+(l,m)\delta $ \end{enumerate} \item Set $\tilde{\bm{u}}=T\bm{u}+\bm{x}$ using enough digits of precision such that the numerical error is at most $|\bm{\epsilon}_{2}|$, where $\bm{u}\equiv\bm{u}_{1}\oplus\dots\oplus\bm{u}_{n}$ and $\tilde{\bm{u}}\equiv\tilde{\bm{u}}_{1}\oplus\dots\oplus\tilde{\bm{u}}_{n}$.\\ (This step corresponds to updating the sampled state according to the linear optical transformation.) \item For each $j=1,\ldots, n$ execute a through e. (This step is to simulate drawing a measurement outcome from the Gaussian measurement distribution centered at $\tilde{\bm{u}}$.) \begin{enumerate} \item For each integer $l,m \in \left[-\frac{\sqrt{|\mathcal{A}|}}{\delta},\frac{\sqrt{|\mathcal{A}|}}{\delta}\right]$ set $\text{Pr} _{\text{{\rm sim}},\measj{j}}(l,m):=\exp\left(-\delta(l,m)^{T}V_{\mathcal{M},j}^{-1}\delta(l,m))\right))\cdot\delta^{2}.$ \\(The outcome space is truncated to a size $|\mathcal{A}|$ and discretized into boxes of size $\delta$. This sets a pdf over the centers of the boxes.) \item Set $\text{Pr} _{\text{{\rm sim}},\measj{j}}(l,m)=\text{Pr} _{\text{{\rm sim}},\measj{j}}(l,m)/\sum_{l,m}\text{Pr} _{\text{{\rm sim}},\measj{j}}(l,m).$\\ (This step normalizes the pdf.) \item Draw a sample $(l,m)$ from the pdf $\text{Pr} _{\text{{\rm sim}},\measj{j}}(l,m)$. \item Find integers $(r,s)$ such that $|\delta(l,m)-\Gamma(r,s))|_{\infty}\le\Gamma/2$.\\ (This just amounts to rebinning the outcome distribution into hypercubes of sidelength $\Gamma$; this introduces no errors but does require $\delta\le\Gamma$.) \item Set measurement outcome $\bm{k}_{j}=\tilde{\bm{u}}_j+\Gamma(r,s)$. \end{enumerate} \item Return measurement outcome $\bm{k}\equiv\bm{k}_{1}\oplus\dots\oplus\bm{k}_{n}$. \end{enumerate} \end{algorithm} Our simulation protocol can necessarily only sample from a discrete distribution so we must introduce some notion of how a discrete distribution can be close to the continuous probability density given by the Born rule. The most natural way to do this is to discretize the outcome distribution into boxes of side length $\Gamma$ according to, \begin{definition} Let $\mathcal{N}_{\Gamma}(\bm{k})\subset\mathbb{R}^{2n}$ be a hypercube in outcome space with side length $\Gamma$ centered at the point $\bm{k}$. We define the $\Gamma$ discretization of the quantum outcome distribution to be $\text{Pr} (\bm{K}_{\text{quant,\ensuremath{\Gamma}}}=\bm{k})\equiv\text{Pr} _{{\rm quant},\measz}(\bm{k})\equiv\int_{\mathcal{N}_{\Gamma}(\bm{k})}p(\bm{K}_{\text{quant}}=\tilde{\bm{k}})d\tilde{\bm{k}}$. We can now fix a $\Gamma$ according to our operational requirements for the simulation and ask how well a simulation protocol samples from this distribution. Notice this is an \emph{unavoidable} consequence of trying to approximate a continuous quantity with a discrete system. With this in hand we can give a precise classical simulation protocol by discretizing our naive algorithm, which results in the family of protocols described in algorithm \ref{alg:advanced-Equivalent-Classical-Circuit}. \end{definition} It now easy to see that both the cost of the simulation and the error in our sampling will be a function of the discretization parameters $\delta(n,\epsilon)$ and $|\mathcal{A}(n,\epsilon)|$. If we can pick these parameters such that for fixed error our simulation scheme scales as ${\rm poly}(n)$ then the simulation is efficient. Concretely, \begin{definition} Let $\text{Pr} _{{\rm sim}, \measz}(\bm{k})$ and $\text{Pr} _{{\rm quant},\measz}(\bm{k})$ be the simulated and actual probabilities of obtaining a measured value that is inside a hypercube of volume $\Gamma^{2n}$ centered at a point in phase space $\bm{k}\in \mathbb{R}^{2n}$ for the separable Gaussian measurement $\mathcal{M}(V_{\mathcal{M}})$. A simulation algorithm is efficient if for inputs $n,\Gamma$ and $\epsilon$ there exists a choice of $|\mathcal{A}|,\delta$ such that: \begin{enumerate} \item The 1-norm distance between the $\Gamma$-discretized quantum distribution and the simulator distribution is at most $\epsilon$, \[ |\text{Pr} _{{\rm quant},\measz}(\bm{k})-\text{Pr} _{{\rm sim}, \measz}(\bm{k})|_{1}\le\epsilon \] where we take $\text{Pr} _{{\rm sim}, \measz}(\bm{k})=0$ whenever $\bm{k}$ is outside the domain of $\text{Pr} _{{\rm sim}, \measz}(\bm{k})$ defined by algorithm \ref{alg:advanced-Equivalent-Classical-Circuit}. \item The computational complexity of the simulation scales as ${\rm poly}(n/\epsilon)$. \end{enumerate} \end{definition} We now can show that the simulation of sufficiently smooth separable positive Wigner functions under linear optical operations and Gaussian measurements is efficient. This result is formally stated in the following theorem and proof is given in appendix~\ref{appendix:thmproof}. \begin{theorem}\label{thm:mainthm} The output of algorithm~\ref{alg:advanced-Equivalent-Classical-Circuit} satisfies $|\text{Pr} _{{\rm quant},\measz}(\bm{k})-\text{Pr} _{{\rm sim}, \measz}(\bm{k})|_{1}\le\epsilon$ for input\\ $(\{\bm{\mu}_{\rho_j}\},\{V_{\rho_j}\},\{V_{\mathcal{M},j}\},T,x,\delta,|\bm{\epsilon}_2|,\Gamma,|\mathcal{A}|)$ if \begin{enumerate} \item $n$, $\max_j\{|\bm{\mu}_{\rho_j}|,\|V_{\rho_j},\|V_{\mathcal{M},j}\|\},\|T\|$ and $\|x\|$ are bounded, \item There exist finite $\beta$, $\Lambda$ such that $|\nabla W_{\rho}(\bm{u})|\le n\beta/|\mathcal{A}|^{n}$ and $\left|\nabla_{\bm{k}}M_{\mathcal{M}(V_{\mathcal{M}})}(\bm{k}|{\bm{u}})\right|\le n\Lambda/|\mathcal{A}|^{n}$\label{item:betagammabd} for $\bm{k}\in\mathbb{R}^{2n},\ \bm{u}\in\mathbb{R}^{2n}$, \item $\delta\le\min\left\{\frac{\epsilon}{16\left[\left(1+\|T\|\right)\Lambda+\beta\right]n\sqrt{2n}},\Gamma\right\}$,\label{item:deltabound} \item $\epsilon< 1$, $|\bm{\epsilon}_2|\le \|T\|\delta\sqrt{\frac{n}{2}}$, \item $|\mathcal{A}|\ge 16n\max_{i,j}\left(\left[V_{\rho_i}\right]_{11}+\left[V_{\rho_i}\right]_{22}+\left[V_{\mathcal{M},j}\right]_{11}+\left[V_{\mathcal{M},j}\right]_{22}\right)/\epsilon$,\label{item:areabound} \item The finite precision error from each oracle $\mathcal{W}_{\rho_j,\eta}$ satisfies $\eta\le\frac{\epsilon}{8n|\mathcal{A}|^{n}}$. \end{enumerate} Furthermore, if we assign unit cost to evaluations of $\mathcal{W}_{\rho_j,\eta}$ and Gaussian functions unit cost then the computational complexity of the algorithm is $O\left(n^{5}\left(\max_i\|V_{\rho_i}\|+\max_j\|V_{\mathcal{M},j}\|\right)\left[\Lambda^2\|T\|^2+\beta^2\right]/\epsilon^3\right),$ which implies efficiency. \label{thm:main} \end{theorem} The key insight of this theorem is that the assumption of Gaussian preparations made in the continuous variable Gottesman--Knill theorem can be relaxed~\cite{Bartlett2002Efficient,Bartlett2002Efficient2}, and that a much wider class of quantum dynamics can be efficiently simulated than previously thought. Indeed, although we stated the algorithm only for product state inputs and product measurements we can now see that this restriction was unnecessary. In fact, our simulation scheme works for any positively represented input and measurement as long as it is possible to efficiently sample from the corresponding distribution. The product assumption is a sufficient but not necessary condition for this efficient sampling. We also note that the algorithm requires us to know the mean and covariance matrices of the distributions, which might be hard to compute analytically. However, since we already require efficient sampling we may appeal to Monte Carlo estimation protocols to determine these quantities within acceptable error tolerances. This extension of the continuous variable Gottesman-Knill theorem places much stronger limitations on the input states that can be used for continuous variable quantum computation and underscores the significance of negativity in the Wigner function as a resource for quantum computation (in analogy to recent results for discrete systems). In particular, we will show that Theorem~\ref{thm:mainthm} places a minimum efficiency for a class of photonic thermal states beyond which the states cannot be used as a resource for linear optical quantum computation with Gaussian measurements. \section{Efficient Simulation of Single Photon--Added--Thermal States}\label{sec:SPATS} The debate over the {}``correct'' definition of classicality for quantum states of light has been a long and, at times, fierce one. One of the most common notions of classicality is whether a state can be represented as a convex combination of Gaussian states. Here we have shown that, in the context of computational power, such a condition is superseded by the condition of positivity of the Wigner function. In this section we give a concrete example of an interesting class of states which are not Gaussian but which have positive Wigner representation and thereby admit an efficient classical simulation. We consider the experimentally accessible class of states called single-photon-added thermal states (SPATS) \cite{Agarwal1992Nonclassical,Zavatta2004QuantumtoClassical,Parigi2007Probing}: \[ \rho(\overline{n})=\frac{1}{\overline{n}(\overline{n}+1)}\sum_{n=0}^{\infty}\left(\frac{\overline{n}}{\overline{n}+1}\right)^{n}n\op{n}{n}, \] where $\overline{n}$ is the mean photon number---given by, in terms of temperature $T$, the Planck distribution $\overline{n}=1/(\exp(1/T)-1)$. It is known that all states in this class are outside the convex hull of Gaussian states and have negative Wigner function for finite temperatures. Under experimentally realistic conditions we must consider states subjected to losses. In general, losses can be modeled as an interaction with a vacuum mode at a beam-splitter with transmittance $\eta$, also called the \emph{quantum efficiency}. The loss rate is then $1-\eta$. The Wigner function of the SPATS after this channel, which we call limited-efficiency SPATS (or LESPATS for short) is \cite{Zavatta2007Experimental} \[ W_{\rho(\overline{n},\eta)}(q,p)=\frac{2}{\pi}\frac{1+2\eta[\overline{n}+2(\overline{n}+1)(q^{2}+p^{2})-2\overline{n}\eta-1]}{(1+2\overline{n}\eta)^{3}}\exp\left(-\frac{2(q^{2}+p^{2})}{1+2\overline{n}\eta}\right). \] Note that the most negative value of the LESPATS is at $(q,p)=(0,0)$ for all $\eta$ and $\overline{n}$. Thus, we consider the quantity \[ W_{\rho(\overline{n},\eta)}(0,0)=\frac{2}{\pi}\frac{1+2\eta(\overline{n}-2\overline{n}\eta-1)}{(1+2\overline{n}\eta)^{3}}. \] By inspection, we can see that for efficiencies of $\eta\leq0.5$, the Wigner function of the LESPATS is positive $W_{\rho(\overline{n},\eta)}(q,p)>0$. Thus, efficiencies of $\eta>0.5$ are necessary for quantum computational speed-up with SPATS. Note however that for $\eta\leq0.5$, the LESPATS are not inside the convex hull of Gaussian states. To see this most clearly, we require a different phase space distribution. The Glauber $P$-function (see for example \cite{Schleich2001Quantum}) is defined as \[ \rho=\iint P_{\rho}(q,p)\op\alpha\alpha dqdp, \] where $\ket{\alpha}$ are the \emph{coherent states}, which are vacuum states that have been displaced in phase space (symmetric Gaussian states). Note that if $P_{\rho}$ is a probability distribution then $\rho$ is in the convex hull of coherent states. The $P$-function of the LESPATS is \cite{Agarwal1992Nonclassical} \[ P_{\rho(\overline{n},\eta)}(q,p)=\frac{1}{\pi\overline{n}^{3}\eta}\left[(\overline{n}+1)\frac{q^{2}+p^{2}}{\eta}-\overline{n}\right]\exp\left(-\frac{q^{2}+p^{2}}{\overline{n}\eta}\right). \] Note that this function is negative for all allowed values of $\eta$ and $\overline{n}$. Thus, the LESPATS are always outsides the convex hull of Gaussian states but are bound universal states \cite{Veitch2012Negative} for $\eta\leq0.5$. To illustrate this, we compare the negativity of the Wigner function with the distance to the convex hull of Gaussian states. The distance we use is based on fidelity, which can be computed using phase space distributions as \[ F(\rho(\overline{n},\eta),\op00)=\text{Tr}[\rho(\overline{n},\eta)\op00]=\pi\iint P_{\rho(\overline{n},\eta)}(q,p)Q_{\op00}(q,p)dqdp, \] where the $Q$-function \[ Q_{\op00}(q,p)=\frac{1}{\pi}\exp(-(q^{2}+p^{2})) \] is \emph{dual} to the $P$-function% \footnote{This duality relationship holds for any phase space representation \cite{Ferrie2011Quasiprobability}.% }. Using a standard table of Gaussian integrals we find \[ F(\rho(\overline{n},\eta),\op00)=\frac{1-\eta}{(1+\overline{n}\eta)^{2}}. \] This effect is demonstrated in figure \ref{fig:SPATS}. Note that, for any state, a quantum efficiency of $\eta\leq 0.5$ is sufficient to ensure membership of the convex hull of states that have positive Wigner representation. This effect is mirrored in the discrete case \cite{Veitch2012Negative}, where depolarizing noise of 50\% is sufficient to ensure membership of the polytope of states with positive \emph{discrete} Wigner function, when starting from any qudit state. \begin{figure}[t] \centering{}{\includegraphics[width=0.9\linewidth]{SPATS3}} \caption{\label{fig:SPATS} Negativity of Wigner function on the left and fidelity to vacuum (``distance'' to convex hull of Gaussians) on the right for varying $\bar{n}$ and $\eta$ (note: $\eta=1$ corresponds to no losses). In both figures ``+'' indicates the region of non-negative states and ``-'' indicates the region of states with negative Wigner function. The region of non-negative states ($\eta\leq0.5$) is the region of bound universal states. This is clear as the Wigner function is positive yet the states still lie outside the convex hull of coherent states since the $P$-function is always negative. Notice that the fidelity distance to the convex hull (as measured by the fidelity to the nearest state, $\ket{0}$) is significantly less than 1, suggesting that the region of bound states is quite large.} \end{figure} \section{Conclusion}\label{sec:conclusion} We have shown that Gaussian quantum computations utilizing separable initial preparations with positive Wigner function are classically efficiently simulable. Since such states lie outside the convex hull of Gaussian states, we have identified a large class of \emph{bound states}: states that cannot be prepared using Gaussian operations, yet do not permit universal quantum computation. We illustrated this class using the example of single-photon-added-thermal-states, showing that quantum efficiencies of 50\% are necessary for quantum computation. Effort has been extended beyond qualitatively defining negativity as quantumness to \emph{quantifying} quantumness via negativity. In terms of the Wigner function, the \emph{volume} of the negative parts of the represented quantum state has been suggested as the appropriate measure of quantumness \cite{Kenfack2004Negativity}. The \emph{distance} (in some some preferred norm on the space of Hermitian operators) to the convex subset of positive Wigner functions was suggested to quantify quantumness in reference \cite{Mari2010Directly}. The \emph{volume} of negativity of the Wigner function (and, in the finite dimensional case, the sum of the negative values) is a ``good'' measure of non-classicality since it is monotonic under Gaussian operators; that is, Gaussian operations cannot increase the volume of the negative regions in phase space. Reference \cite{Bartlett2003Requirement} nicely summarized what was known at the time about continuous variable quantum computation. The table presented there is reproduced below in table \ref{table:review} with some more recent results. The field began with Lloyd and Braunstein's observation that non-linear optical processes are sufficient for universal continuous variable quantum computation. Later, it was shown for discrete variable encodings that linear optics is sufficient provided photon counting measurements are available \cite{Knill2001Scheme,Gottesman2001Encoding}. The continuous variable analog of the \emph{measurement-based} model shows that preparation of single photon state preparation is also sufficient \cite{Menicucci2006Universal}. More recently, the result of Aaronson and Arkhipov \cite{Aaronson2010Computational} shows that preparing and measuring single photon states (without post-selection) is equivalent to a sampling problem that is thought to be hard classically---but it still manages to (probably) not be universal for quantum computation. It is possible that the Aaronson and Arkhipov model may be intermediately between classically efficiently simulatable and universal for quantum computation. Another suspected model of this type is the {}``one-clean-qubit'' model of Knill and Laflamme \cite{Knill1998Power}. The key point for this latter model is that uses highly mixed states. Mixed states have not been given full consideration for continuous variable quantum computation. Here we have shown, via the Wigner phase space formalism and independent of purity, negative representation is necessary for universal quantum computation. Moreover, any computation that uses states possessing a positive Wigner function is classically efficiently simulatable. It would be quite interesting if this condition turned out also to be sufficient as this would provide a sharp boundary between quantum and classical systems with regard to their computational power. \begin{acknowledgements} We thank Earl Campbell and Christian Weedbrook for helpful comments. The authors acknowledge financial support from the Government of Canada through NSERC, CIFAR, USARO-DTO. After completion of this work, we were made aware of \cite{Mari2012Positive}, who obtain a similar result for a different (namely, local) set of dynamical transformations. \end{acknowledgements}
\section{Introduction}\label{intro} Multipartite states are an important resource for many areas of quantum information. Understanding the features that give rise to their usefulness is still a question under much investigation. Entanglement has become recognized as a key feature. However, the question becomes involved in the multipartite settings with different classes of entanglement~\cite{RevModPhys.81.865}, having potentially different roles in recognizing good resources. Intimately related to entanglement is notion of nonlocality~\cite{PhysRev.47.777,bell1964einstein}, though it is known they are not the same~\cite{PhysRevA.40.4277}. Very little is currently known of the richness of multipartite state space and how, if at all, it is exhibited through nonlocality as it is through entanglement theory, with some recent progress in this direction, for example~\cite{PhysRevLett.108.110501,PhysRevLett.108.210407}. Amongst multipartite states, graph states and symmetric states dominate experimental progress, with experiments with up to 10 qubits~\cite{Lu:2007fk,PhysRevLett.103.020504,PhysRevLett.103.020503,Gao:2010uq}. These two classes represent potentially very different resources for quantum information and have different entanglement features. By exploiting their entanglement properties~\cite{casati2006quantum}, graph states are useful for many tasks such as error correction~\cite{PhysRevA.69.062311}, measurement based quantum computation (MBQC)~\cite{PhysRevLett.86.5188,danos2007measurement,browne2007generalized} and secret sharing~\cite{PhysRevA.78.042309}. In addition to the entanglement properties, nonlocal features of graph states have also been well-studied~\cite{PhysRevLett.95.120405}. Most of these properties are studied via an elegant mathematical tool: the stabilizer formalism~\cite{gottesmanthesis}. On the other hand, permutation symmetric states occur very often in optics, in the form of Dicke states~\cite{PhysRevLett.103.020504,PhysRevLett.103.020503,PhysRevA.81.032316} and in many-body physics as ground states for example in some Bose-Hubbard models. Similar to graph states, the study of permutation symmetric states can be carried out in an elegant mathematical framework as well: the Majorana representation~\cite{majorana1932atomi}, where symmetric states of $n$ qubits can be represented as $n$ points on the surface of a sphere. In terms of multiparty entanglement properties, much has been learnt using this representation ~\cite{PhysRevLett.103.070503,PhysRevA.81.052315,aulbach2011symmetric,springerlink,PhysRevA.83.042332,PhysRevLett.106.180502,martinthesis}. The nonlocality of symmetric states has been studied recently also by using the Majorana representation~\cite{PhysRevLett.108.210407}, where it was shown that all symmetric states can violate a Bell inequality (it has very recently been shown that all entangled pure states can violate the same inequality~\cite{PhysRevLett.109.120402}). It was also shown in~\cite{PhysRevLett.108.210407} that the degeneracy of the points in the Majorana representation (i.e. when points sit on top of each other) gives persistency of correlations to sets of subsystems. Since degeneracy of Majorana points also separates entanglement classes~\cite{PhysRevLett.103.070503}, this indicates a connection between entanglement classes and nonlocal properties. As well as the general interest in exploring the texture of multipartite state space, there is some practical interest in understanding the relationship between entanglement and nonlocality. Using the entanglement or nonlocal properties of multipartite states in the real world poses many experimental challenges. Unavoidable experimental inaccuracies like misalignment, noise and detector inefficiencies can render the outcome of an experiment meaningless. In quantum cryptography, for example, the presence of noise and detector inefficiencies can mask effective effective attacks on the security of the key distribution protocol~\cite{q2007journal,1367-2630-12-11-113026,Lydersen2010fk}. In entanglement theory, misalignment when trying to witness entanglement can lead to mistaken claims of the existence of entanglement~\cite{PhysRevLett.106.250404}. One solution to these problems is to make tangible claims without any assumptions about the measurement device, hence the name \emph{device independent}. There is a natural connection to discussions of nonlocality since Bell type arguments do not rely on any statements about measurements, only their statistics. Using these ideas, device independent proofs and tests have already been used extensively in quantum cryptography and secure communications~\cite{PhysRevLett.98.230501,1367-2630-11-4-045021,PhysRevLett.105.230501,1751-8121-44-9-095305,masanes2011secure,Pironio2010fk}, and device independent entanglement witnesses~\cite{PhysRevLett.106.250404} have been proposed. Recent results have shown device independent tests which are able to discriminate states that are inequivalent under local unitaries and permutation of systems (LUP)~\cite{PhysRevLett.108.110501}. In this work we further the study of nonlocality of symmetric states using the inequalities and techniques raised in~\cite{PhysRevLett.108.210407}, to study deeper how the nonlocality exposed is related to entanglement classes and the usefulness of the states. We will offer new evidence of connection between the nonlocal properties of states and their entanglement properties through device independent classification of states via violation of Bell inequalities presented in ~\cite{PhysRevLett.108.210407}. We then look at how the violation of inequalities scales with the number of systems. We see that violation is upper bounded by entanglement because of the form of the inequality, in particular that it has only one positive term. Related to this we also look at what can be said about the monogamy of the correlations that can be witnessed (a useful property for quantum cryptography~\cite{PhysRevLett.97.170409,PhysRevLett.108.100401}).We see here again that the inequalities of~\cite{PhysRevLett.108.210407} are not suited for showing strict monogamy. Motivated by this we then introduce new inequalities with more positive terms which show good scaling of violation with $n$, implying monogamy in the high $n$ limit. This paper is organized as follows. In section~\ref{bg}, we give a brief introduction to some concepts and results which we will use in later sections to make the paper self-contained. We recall the main results of~\cite{PhysRevLett.108.210407}: the inequalities to show the nonlocality of all symmetric states and the procedure to find measurement bases to violate them for almost all symmetric states. Subsequently, in section~\ref{sdp}, we introduce the semidefinite programming (SDP) techniques we use to obtain numerical results, with an example showing the violations of the inequalities introduced in section~\ref{bg} for a class of states. Then in section~\ref{mobius} we use SDP to show how these inequalities allow us to have a device independent discrimination of multipartite entangled state classes, and we see further evidence that degeneracy of Majorana points leads to natural classification with respect to nonlocality. Section~\ref{analytical} shows how the violations of these inequalities scale in the case of large $n$ for two most common sets of symmetric states: the W states and the GHZ states. In section~\ref{monogamy}, we discuss monogamy of entanglement and monogamy of correlations - that is, how much entanglement and correlations can be shared. We see how the inequalities in~\cite{PhysRevLett.108.210407} are not suited to showing strict monogamy, though bounds on how much correlations can be shared can be derived using the methods of~\cite{PhysRevLett.102.030403}. We then generalize a recently presented inequality for W states~\cite{1367-2630-13-5-053054} to all Dicke states and show violation limits to maximal for large $n$ and the correlations from these nonlocal tests are strictly monogamous when the number of parties goes to infinity. We finish with discussions and conclusion. \section{Background}\label{bg} We start by giving some background, introducing notation, presenting the inequalities in~\cite{PhysRevLett.108.210407} and how to find the measurement settings to violate them. We consider two dichotomic measurement settings per party, and we use $0$ and $1$ to label both the settings and the outcomes. For measurement settings $M_1, ..M_n$ we denote the probability of getting results $m_1, ...m_n$ as $P(m_1, ...m_n|M_1,...,M_n)$. For example, $P(000|111)$ gives the probability that all three parties obtain result $0$ having measured in setting $1$. The inequality, which we call $\mathcal{P}^n$, is given by \begin{align} \mathcal{P}^n:= &P(00\ldots00|00\ldots00)\nonumber\\ -&P(00\ldots00|00\ldots01) \nonumber\\ &\vdots\nonumber\\ -&P(00\ldots00|10\ldots00)\nonumber\\ -&P(11\ldots11|11\ldots11)\leq0, \end{align} which must be satisfied by all local hidden variable (LHV) theories. This can be seen since all LHV distributions can be considered as probabilistic mixtures of deterministic local strategies - i.e. ones where $P(m_1, ...m_n|M_1,...,M_n)=\prod_{i}P(m_i|M_i)$ with $P(m_i|M_i)=0$ or $P(m_i|M_i)=1$ - so it is enough to consider these alone~\cite{PhysRevA.64.032112}. Since there is only one positive term in $\mathcal{P}^n$, to have anything greater than zero requires this term to be one. It can easily be seen that this implies that at least one of the negative terms is also one, which gives the desired bound. Following~\cite{PhysRevLett.108.210407}, to show the violation of $\mathcal{P}^n$ for almost all symmetric states, first we note that all symmetric states of $n$ parties can be written as the sum of permutations of $n$ individual qubits in the Majorana representation ~\cite{majorana1932atomi,RevModPhys.17.237}: \begin{align} \ket{\psi}=K\sum_{perm}\ket{\eta_1\ldots\eta_n}\label{decomp}, \end{align} with the Bloch sphere representation of each qubit state $\ket{\eta_i}$ called a \emph{Majorana point} (MP). We use the notation $\ket{\eta_i^{\perp}}$ to indicate the orthogonal state corresponding to the antipodal point of $\ket{\eta_i}$, such that \begin{align} \ip{\eta_i^{\perp}}{\eta_i}=0. \end{align} The MPs are then used to find a suitable basis measurements in which show a violation of $\mathcal{P}^n$. In the prescription described in~\cite{PhysRevLett.108.210407}, for each party, the setting $1$ is chosen to be the basis defined by one of the MPs, say $\ket{\eta_i}$ (associated to outcome $0$) and its orthogonal state $\ket{\eta_i^{\perp}}$ (associated to outcome $1$). Then it is easy to see that \begin{align} P(11\ldots11|11\ldots11)=|(\bra{\eta_i^{\perp}})^{\otimes n}\ket{\psi}|^2=0. \end{align} To find the basis corresponding to setting $0$, we first notice that the $n-1$ party state \begin{align} \ket{\psi'}=\ip{\eta_i}{\psi} \end{align} is also a symmetric state, so we can use the same idea. Denoting the $0$ outcome of the basis $0$ by $|0\rangle$, we have \begin{align} P(00\ldots00|00\ldots01)&=|(\bra{0})^{\otimes {n-1}}\ip{\eta_i}{\psi}|^2 \nonumber \\ &= |(\bra{0})^{\otimes {n-1}}\ket{\psi'}|^2. \end{align} We can then use the Majorana points of $\ket{\psi'}$ for the $0$ basis, as above, to take probabilities $P(00\ldots00|00\ldots01)$ to $P(00\ldots00|10\ldots00)$ zero. Proposition~1 in~\cite{PhysRevLett.108.210407} guarantees that there exists such a choice which also makes $P(00\ldots00|00\ldots00)>0$ for all symmetric states except Dicke states. For Dicke states \begin{align} \ket{S(n,k)}=\frac{1}{\sqrt{n \choose k}}(\sum_{perm}\ket{\underbrace{0\ldots0}_{n-k}\underbrace{1\ldots1}_{k}}), \end{align} this procedure no longer applies, but $\mathcal{P}^n$ can still be violated by another basis choice (also found in~\cite{PhysRevLett.108.210407}). In the case that not all $\ket{\eta_i}$ are distinct, we say the state $\ket{\psi}$ is \emph{degenerate}. If $d$ MPs sit on top of each other, we say there is degeneracy $d$. We can extend $\mathcal{P}^n$ to reflect the degeneracy, defining the extended inequality as: \begin{align} \label{def: degQ} \mathcal{Q}_{d}^n := &\mathcal{P}^n - P(\underbrace{11\ldots1}_{n-1}|\underbrace{11\ldots1}_{n-1}) - ... - P(\underbrace{11\ldots1}_{n-d+1}|\underbrace{11\ldots1}_{n-d+1})\nonumber\\ & \leq 0. \end{align} To calculate the new probabilities, we trace out one party for each term (because of the symmetry of the state, it does not matter which party we trace out): \begin{align} P(\underbrace{11\ldots1}_{n-1}|\underbrace{11\ldots1}_{n-1})&=Tr(\rho_1 \underbrace{M_2^{00}\otimes\ldots\otimes M_n^{00}}_{n-1})\nonumber\\ &\vdots\nonumber\\ P(\underbrace{11\ldots1}_{n-d+1}|\underbrace{11\ldots1}_{n-d+1})&=Tr(\rho_{d-1} \underbrace{M_d^{00}\otimes\ldots\otimes M_n^{00}}_{n-d+1}), \end{align} where $\rho_1=Tr_i(\ket{\psi}\bra{\psi})$, $\rho_2=Tr_j(\rho_1)$, etc. It can easily be seen that if the MP chosen for basis $1$ has degeneracy $d$, then these probabilities will be zero, hence, the same measurements will lead to a violation of $\mathcal{Q}_{d}^n$. In this sense the persistency of the correlations to subsystems is guaranteed by the degeneracy of MPs. The connection with degeneracy makes a connection to entanglement classes. Since the degeneracy of points is something that cannot change under SLOCC, states with different degeneracy belong to different classes~\cite{PhysRevLett.103.070503}. As noted in~\cite{PhysRevLett.108.210407}, the above discussion has an interpretation that having at least one MP with degeneracy $d$ automatically means it is possible to violate $\mathcal{Q}_{d}^n$, giving some kind of operational significance to the class. We will study this further in sections III and IV to understand this more. We will also be interested in what the maximum violation of these inequalities can be. With respect to this question there is a straightforward, though perhaps surprising, bound given be the geometric measure of entanglement~\cite{PhysRevA.80.032324} \begin{eqnarray} \label{Eqn: Def Eg} E_g(|\psi\rangle) = \min_{|\Phi\rangle \in {\rm Pro}} -\log_2 (|\langle \Phi | \psi \rangle |^2), \end{eqnarray} where ${\rm Pro}$ is the set of product states. It is apparent from the definitions above that for a state $\ket{\psi}$ with entanglement $E_g$, the violations of both $\mathcal{P}^n$ and $\mathcal{Q}_{d}^n$ are bounded by $P(00\ldots00|00\ldots00)$, which in turn is bounded by $\frac{1}{2^{E_g}}$. That is, \begin{align} &\mathcal{P}^n\leq \frac{1}{2^{E_g}}&\mathcal{Q}_d^n\leq \frac{1}{2^{E_g}} \end{align} Thus, states with very high entanglement necessarily violate at best by a small amount. The geometric measure is easy to calculate for symmetric states~\cite{PhysRevA.80.032324,aulbach2010maximally,springerlink,PhysRevA.83.042332,martinthesis}, with their entanglement properties with respect to the geometric measure relatively well-known~\cite{PhysRevA.83.042332,springerlink,martinthesis}. However, as we will see in later sections (Figures~\ref{000t_plot} and~\ref{ghz_w_plot}), the geometric measure does not always give a good bound on the violations of $\mathcal{P}^n$ and $\mathcal{Q}_{d}^n$, because of the presence of many negative terms in their expressions. In particular this is true with scaling in number of parties, $n$, where we see that violation decreases as entanglement increases in section~\ref{analytical}. This will later have implications on what can be said about monogamy in section~\ref{monogamy}. The bound from the entanglement will later motivate the introduction of new inequalities with more positive terms so that large violation is possible. \section{Semidefinite programming techniques}\label{sdp} In this section we introduce the techniques used to find numerical bounds on the violation of $\mathcal{P}^n$ and $\mathcal{Q}_{d}^n$ for given states. These will, in turn, be used to look at trends in violation with degeneracy, and also to show device independent separation of state classes. Semidefinite programming was developed in the 90s as a tool to study convex optimization problems~\cite{vandenberghe1996semidefinite}. The method has been adapted in early 2000s as a way to numerically find the global extrema of a real-valued polynomial~\cite{lasserre2001global}. Also around this time, the study of multiparty nonlocality produced increasingly complex results, which made it hard to obtain analytical properties about various multiparty Bell inequalities. As a result, numerical studies about the optimality and violations of these inequalities began to emerge~\cite{PhysRevA.64.032112}~\cite{PhysRevA.64.014102}. In 2006, Wehner~\cite{PhysRevA.73.022110} used SDP as an analytical tool to both prove the original Tsirelson bound for the CHSH inequality and to find new bounds for the generalized CHSH inequality with $n$ settings and 2 outcomes per setting. Since then, SDP has been employed as a numerical tool to study various aspects of multiparty entanglement and features of multiparty nonlocality, for example in ~\cite{PhysRevLett.98.010401}~\cite{PhysRevLett.100.210503}. A recent paper~\cite{PhysRevLett.108.110501} used SDP to show that one can distinguish two different classes of entangled states based on violations of Bell inequalities. For our purposes, we employ a similar technique to the one used in~\cite{PhysRevLett.108.110501}. Since, without loss of generality, we only use projective measurements~\cite{PhysRevA.64.032112} and probabilities instead of expectation values, the measurement operator we use is different. Suppose each player $i$ can measure either one of two bases and obtain either one of two possible outcomes. We model these four different situations by four measurement operators: \begin{align} M_i^{00}&=\frac{1}{2}(\mathbb{I}_i+\alpha_{i0}\mathcal{X}_i+\beta_{i0}\mathcal{Y}_i+\gamma_{i0}\mathcal{Z}_i)\nonumber\\ M_i^{01}&=\frac{1}{2}(\mathbb{I}_i-\alpha_{i0}\mathcal{X}_i-\beta_{i0}\mathcal{Y}_i-\gamma_{i0}\mathcal{Z}_i)\nonumber\\ M_i^{10}&=\frac{1}{2}(\mathbb{I}_i+\alpha_{i1}\mathcal{X}_i+\beta_{i1}\mathcal{Y}_i+\gamma_{i1}\mathcal{Z}_i)\nonumber\\ M_i^{11}&=\frac{1}{2}(\mathbb{I}_i-\alpha_{i1}\mathcal{X}_i-\beta_{i1}\mathcal{Y}_i-\gamma_{i1}\mathcal{Z}_i), \end{align} where $M_i^{jk}$ denotes the player $i$ chooses to measure in basis $j$ and obtains the outcome $k$, and $v_i^0=(\alpha_{i0},\beta_{i0},\gamma_{i0})$, $v_i^1=(\alpha_{i1},\beta_{i1},\gamma_{i1})$ are two unit vectors in $\mathbb{R}^3$. Now we can write the probabilities in $\mathcal{P}_n$ using these single-qubit measurement operators: \begin{align} P(0\ldots0|0\ldots0)&=Tr(\rho M_1^{00}\otimes\ldots\otimes M_n^{00})\nonumber\\ P(0\ldots0|0\ldots1)&=Tr(\rho M_1^{00}\otimes\ldots\otimes M_n^{10})\nonumber\\ \vdots\nonumber\\ P(0\ldots0|1\ldots0)&=Tr(\rho M_1^{10}\otimes\ldots\otimes M_n^{00})\nonumber\\ P(1\ldots1|1\ldots1)&=Tr(\rho M_1^{11}\otimes\ldots\otimes M_n^{11}), \end{align} where $\rho=\ket{\psi}\bra{\psi}$ is the density matrix of a $n$-qubit permutation symmetric state $\ket{\psi}$. Rewriting $\mathcal{P}_n$ this way results in a vector polynomial of $2n$ variables ($v_i^0$ and $v_i^1$ for each $i$) \begin{align} &\mathcal{V}(v_1^0,v_1^1,\ldots, v_n^0, v_n^1)=\nonumber\\ &Tr(\rho M_1^{00}\otimes\ldots\otimes M_n^{00})\nonumber\\ -&Tr(\rho M_1^{00}\otimes\ldots\otimes M_n^{10})\nonumber\\ \vdots\nonumber\\ -&Tr(\rho M_1^{10}\otimes\ldots\otimes M_n^{00})\nonumber\\ -&Tr(\rho M_1^{11}\otimes\ldots\otimes M_n^{11}). \end{align} The goal of an SDP program is to maximize $\mathcal{V}(v_1^0,v_1^1,\ldots, v_n^0, v_n^1)$, subject to the constraint that the Gram matrix formed by the vectors $v_i^0$ and $v_i^1$ is positive semidefinite~\cite{horn1990matrix}. As a first example of the use of SDP, we compute the violation of $\mathcal{P}^4$ and $\mathcal{Q}^4_3$ by a special set of states: the states $\ket{000\theta}=K\sum_{perm}\ket{0}\otimes\ket{0}\otimes\ket{0}\otimes(cos(\frac{\theta}{2})\ket{0}+sin(\frac{\theta}{2})\ket{1})$, with three MPs at the north pole and the other MP varies from $\ket{0}$ (a product state) to $\ket{1}$ (the 4-party W state). For this set of states, the geometric measure of entanglement is easy to calculate by simply searching product symmetric states~\cite{PhysRevA.80.032324}. It should be noted, however, the high level of degeneracy of the state $\ket{000\theta}$ makes it difficult for the SDP program to compute a good bound. This is probably due to the fact that the polynomial defining the SDP problem has high degeneracy, carried over from the degeneracy of the state. The SDP solver we use, SDPNAL, is known to be inaccurate when the optimal solutions are degenerate~\cite{zhao2010newton}. To ease computation for this example, we assume that every player measures in the same bases, which numerically seems to be a reasonable assumption. When we insist every player measures in the same bases, most of the values computed by SDP and plotted in Fig.~\ref{000t_plot} can be certified numerically, meaning there are quantum measurements which can achieve these values. The results are shown in Fig.~\ref{000t_plot}. Note that in later sections we will not make this assumption (unless stated explicitly), here we do so simply as an example to see trends. \begin{figure}[htbp] \begin{center} \includegraphics[width=230px,keepaspectratio=true]{Fig1.eps} \caption{A comparison of $\frac{1}{2^{Eg}}$ (\ding{108}), the violation of $\mathcal{P}^4$ (\ding{72}) and $\mathcal{Q}^4_3$ (\ding{113}) for the states $\ket{000\theta}$, when $\theta$ varies from $0$ to $\pi$. } \label{000t_plot} \end{center} \end{figure} It can be seen that the violations of $\mathcal{Q}^4_3$ follow very closely the violations of $\mathcal{P}^4$. Note that in the earlier discussion of degeneracy, where we argued that degeneracy guarantees violation of $\mathcal{Q}^n_d$ (which we can understand as correlations persisting to fewer numbers of parties) we were looking at a particular prescription of measurements, which may not be maximal. In these numerics we have searched over all bases (assuming players all measure in the same basis), indicating that the maximum violation of $\mathcal{Q}^n_d$ is also persistent in correlations of subsets of parties. We also notice that the upper bound given by entanglement is closer to the violation as the angle tends towards $\pi$. \section{Device independent classification of states}\label{mobius} Although there are clear connections between the violation of $Q^n_d$ and SLOCC entanglement classes through degeneracy of MPs - degeneracy $d$ guarentees always violation of $Q^n_d$ - the relationship is not as clear as we might like. An immediate question is the one raised above, the violation of $Q^n_d$ is guaranteed by the prescription using the Majorana representation, but what about the maximal violation? Can we say that degeneracy guarantees that the level of violation stays high? Although we no longer have the analytic tools for general violation, we will see that numerics seems to indicate this is the case, at least for W states. A deeper question though is what we can really understand from this. We would really like to know if it is possible to use these ideas and results to separate classes of states - so that different classes can really be differentiated by their nonlocal properties. This would lead to new ways of searching for new applications of states, as well as ways of probing the texture of multipartite states. To answer this, we will first go more into the subtle questions surrounding the classification of states, and then we will see some examples of how some separation of classes can be made. On a practical level, it seems clear that different multipartite entangled states have different entanglement and locality properties. Famously GHZ states are highly nonlocal, but are highly sensitive to loss of systems - losing even one system takes them to a separable (hence `local' state), whereas W states do not have the same extreme nonlocality~\cite{PhysRevA.63.022104}, but losing systems does not destroy the entanglement. In turn, different types of states may have different uses for quantum information. The question of how to classify states in terms of entanglement and locality is a difficult one, particularly when we want to talk about how different `classes' might be meaningful either for different quantum information tasks, or their potential roles in many-body physics. Within entanglement theory, the most standard approach is to define two states as equivalent if they can be mapped to one another using only local operations and classical communication (LOCC), with some non-zero probability. This method of classification leads to what are called SLOCC classes of states (the S standing for Stochasitic)~\cite{PhysRevA.62.062314}. Intuitively this classification is appealing since it separates states which cannot be reached from each other in the distributed setting, even with the aid of classical communication. In terms of how one might classify states with respect to locality, there are several approaches. The standard setting for locality questions is one in which parties are not allowed to communicate classically - at least not after they have been told what bases to measure in, they may do before hand, for example to share classical randomness. Several options arise. In~\cite{PhysRevA.83.022328} it is proposed that a reasonable classification is to consider equivalence under local unitaries and permutation of systems (we denote this LUP). One may also consider states equivalent under local operations, which is in turn equivalent to local unitaries (we donate this LU). When considering correlations alone, without necessarily taking recourse to quantum states, in~\cite{PhysRevLett.109.070401} a classification is presented called wiring and classical communication prior to inputs (WCCPI) - the wiring is essentially the idea of using multiple copies of the resource (which could be a quantum state or `box' giving a certain probability distribution) and allowing different ways of combining them. We do not consider the WCCPI classification further here, and rather focus on single copy classifications. For all the classifications mentioned above, however, several difficulties emerge, which seem to limit their usefulness. First of all, there can be an infinite continuum of classes (for LUP and LU this is already true for two quibits, for SLOCC it is true for four or more~\cite{PhysRevA.62.062314}). Second, and related to this, it is possible to have two states which are arbitrarily close to each other which are in different classes. This means that two states, which behave in almost exactly the same way for all possible experiments, can be in different classes. It is clear then that it is not possible to separate all classes of states in terms of their physical properties and in turn that the physical properties cannot be sensitive to all these classifications. Nevertheless, there does seem to be some difference between states, which can be identified through these classifications. For example, as we saw earlier, states of certain classes guarantee resistance of correlations to loss of systems, for both the LUP~\cite{PhysRevLett.108.110501} and the SLOCC~\cite{PhysRevLett.108.210407} classifications (through the degeneracy of MPs as mentioned earlier). In~\cite{PhysRevLett.108.110501} this was used to separate two LUP classes in a device independent way. Here we will use our inequalities to identify different sets of LUP classes of states of four qubits, hence also, in a device independent way. The LUP and LU classifications are well suited to discriminate via inequality violation because the maximum violation of an inequality is searched for over all measurement bases - which is equivalent to searching over all local unitaries. Thus, if we can say that a particular state cannot violate an inequality more than a certain amount (using SDP techniques for example, as we do here), this means that no state in the same LU class can either. If the state is symmetric it also means no state in the same LUP state can either. The states we choose are also in different SLOCC classes (note, however, that the fact that no LU or LUP equivalent state can violate more than the amount we state does not necessarily mean that there does not exist an SLOCC equivalent state which can). Since this is done via violation of Bell-like inequalities - which makes no recourse to what measurements are made, this classification is done in a device independent way. For the classification, we will consider three states: the tetrahedron state $\ket{T}=\sqrt{\frac{1}{3}}\ket{S(4,0)}+\sqrt{\frac{2}{3}}\ket{S(4,3)}$, the 4-qubit GHZ state $\ket{GHZ_4}=\frac{1}{\sqrt{2}}(\ket{0000}+\ket{1111})$, and the state $\ket{000+}=K\sum_{perm}\ket{000+}=\frac{2}{\sqrt{5}}\ket{0000}+\frac{1}{\sqrt{5}}\ket{S(4,1)}$, which are all SLOCC-inequivalent~\cite{PhysRevA.83.042332,PhysRevLett.103.070503,aulbach2011symmetric,PhysRevLett.106.180502}. We will consider them in two groups: one group consists of $\ket{T}$ and $\ket{000+}$, with differing degeneracy, the other group consists of $\ket{T}$ and $\ket{GHZ_4}$, with the same degeneracy. These are represented in Fig.~\ref{t_000+} and \ref{ghz_t} respectively. We will use numerical maximum violation of $\mathcal{P}^4$ and $\mathcal{Q}^4_3$ obtained from SDP to discriminate the states in a device independent way in each group. The SLOCC-inequivalence of these states can be seen most easily from a recent result~\cite{aulbach2011symmetric}~\cite{PhysRevLett.106.180502}, which has shown that for symmetric states, there is an interesting relationship between SLOCC operations and M\"{o}bius transformations. A \emph{M\"{o}bius transformation} is a function of one variable $z$, defined on the extended complex plane $\mathbb{C}_{\infty}$, which can be written in the form \begin{align} f(z)=\frac{az+b}{cz+d}, \end{align} where $a, b, c ,d$ are complex numbers and $ad-bc\neq0$ (otherwise $f(z)$ is a constant map)~\cite{needhamcomplex}~\cite{beardon2005algebra}. A M\"{o}bius transformation can be seen as a composition of four more elementary steps: translation, complex inversion, expansion and rotation. M\"{o}bius transformations are conformal maps which take circles to circles and preserve the symmetry with respect to circles. The SLOCC-inequivalence of $\ket{T}$, $\ket{GHZ}$ and $\ket{000+}$ can be seen from the fact that it is not possible to change the degeneracy of MPs via M\"{o}bius transformations (see also~\cite{PhysRevLett.103.070503}). When we consider $\ket{T}$ and $\ket{GHZ}$, it is clear from Fig.~\ref{ghz_t} that there is no M\"{o}bius transformation connecting their Majorana points: the MPs of $\ket{GHZ_4}$ all lie on the equator, a M\"{o}bius transformation will map them to another circle, but the MPs of $\ket{T}$ clearly do not form a circle. The equivalence of symmetric states under LU and LUP is given simply by the MP distribution up to rotation of the sphere. This is because any local unitary taking a symmetric state to a symmetric state can be understood as a rotation of the sphere~\cite{PhysRevLett.103.070503,PhysRevA.81.052315} (and that permutation obviously do not change a symmetric state). Thus each of the states we study here are LU and LUP inequivalent. As mentioned, the fact that we search for violation of inequalities over all measurements means that the bounds we present hold for all LU and LUP equivalent states. For the first group, shown in Fig.~\ref{t_000+}, the results are shown in Table~\ref{t_000+_table}. Note that although we do not restrict the measurement bases for $\ket{T}$, as the degeneracy of the state $\ket{000+}$ is very high, we need to restrict the bases to get realistic SDP bounds. \begin{figure}[ht] \centering \subfloat[]{\label{ex:t}\includegraphics[width=120px,keepaspectratio=true]{Fig2a.eps}} \, \subfloat[]{\label{ex:ghz}\includegraphics[width=120px,keepaspectratio=true]{Fig2b.eps}} \caption{The tetrahedron state (a) and the state $\ket{000+}$ (b) in the Majorana representation.} \label{t_000+} \end{figure} \begin{table}[htdp] \begin{center} \begin{tabular}{|c|c|c|} \hline State&$\mathcal{P}^4$&$\mathcal{Q}^4_3$\\ \hline \hline $\ket{T}$&0.1745&-0.0609\\ \hline $\ket{000+}$&0.0142&0.0141\\ \hline \end{tabular} \end{center} \caption{SDP bounds on the maximum violation of $\mathcal{P}^4$ and $\mathcal{Q}^4_3$ for $\ket{T}$ and $\ket{000+}$. Because of computational difficulties the values for $\ket{000+}$ assume that all parties measure in the same basis (numerics indicate this is still optimal). We thus have that a violation of $\mathcal{Q}^4_3$ implies the state is not in the LU class of $|T\rangle$.} \label{t_000+_table} \end{table}% \begin{figure}[ht] \centering \subfloat[]{\label{ex:t}\includegraphics[width=120px,keepaspectratio=true]{Fig3a.eps}} \, \subfloat[]{\label{ex:ghz}\includegraphics[width=120px,keepaspectratio=true]{Fig3b.eps}} \caption{The tetrahedron state (a) and the 4-qubit GHZ state (b) in the Majorana representation.} \label{ghz_t} \end{figure} Table.~\ref{ghz_t_table} shows the bounds for $\mathcal{P}^4$ and $\mathcal{Q}^4_3$ for the second group, shown in Fig.~\ref{ghz_t}, obtained using semidefinite programming techniques described in section~\ref{sdp}, without restricting the measurement bases of parties. \begin{table}[htdp] \begin{center} \begin{tabular}{|c|c|c|} \hline State&$\mathcal{P}^4$&$\mathcal{Q}^4_3$\\ \hline \hline $\ket{T}$&0.1745&-0.0609\\ \hline $\ket{GHZ_4}$&0.1241&0.0563\\ \hline \end{tabular} \end{center} \caption{SDP bounds on the maximum violation of $\mathcal{P}^4$ and $\mathcal{Q}^4_3$ for $\ket{T}$ and $\ket{GHZ_4}$. We thus have that a violation of $\mathcal{P}^4 > 0.1241$ implies the state is not in the LU class of $|GHZ_4\rangle$, and a violation of $\mathcal{Q}^4_3$ implies the state is not in the LU class of $|T\rangle$.} \label{ghz_t_table} \end{table}% From these tables, one can easily envisage device independent tests to discriminate the LUP classes in each group. For the first group, because of the restriction on measurement bases, we have a weaker test. Despite our best numerical checks and the seemingly reasonable assumption on the restriction of measurement of bases, we cannot guarantee that if a state has a violation of $\mathcal{P}^4$ greater than $0.0142$, it is not in the LUP $\ket{000+}$ class. However, we can still conclude that if a state violates $\mathcal{Q}^4_3$ then it cannot be in the $\ket{T}$ class, but must be in the $\ket{000+}$ LUP class. In the second group, if the $\mathcal{P}^4$ test gives a violation $\geq 0.1241$, then the state must not be in the $\ket{GHZ_4}$ LUP class, so must be in the $\ket{T}$ class. Similarly, if the $\mathcal{Q}^4_3$ gives any violation at all, the state cannot be in the $\ket{T}$ LUP class and must be in the $\ket{GHZ_4}$ class. In this case, even though there is no degeneracy, separation can be seen using $\mathcal{Q}^4_3$. \section{Large $n$ results for $\ket{W_n}$ and $\ket{GHZ_n}$}\label{analytical} In this section we study the trends of violations of $\mathcal{P}_n$ for W and GHZ states as $n$ gets large. In terms of monogamy and other applications of nonlocal features (for example communication complexity gains~\cite{RevModPhys.82.665}), we are interested in the value of violation - the higher the better. We are interested then to know how violation scales with $n$. While the use of SDP allows us to study the nonlocality of symmetric states with a few parties, the computational resources required to run the SDP program increase exponentially with the number of parties, which makes it impractical to obtain results for states with more than 4 parties. Luckily, for two commonly studied symmetric states, the W states \begin{align} \ket{W_n}=\ket{S(n,1)}=\frac{1}{\sqrt{n}}(\sum_{perm}\ket{\underbrace{0\ldots 0}_{n-1}1}), \end{align} and the GHZ states \begin{align} \ket{GHZ_n}=\frac{1}{\sqrt{2}}(\ket{\underbrace{0\ldots 0}_{n}}+\ket{\underbrace{1\ldots 1}_{n}}), \end{align} it is possible to calculate analytically the violation of $\mathcal{P}_n$ if the measurement bases are those prescribed in section~\ref{bg} and~\cite{PhysRevLett.108.210407}. This allows us to give bounds on the maximum violation possible and see trends. We will use a combination of this and numerics to approximate the best violation. For the W state, using the bases $\{\ket{+},\ket{-}\}$, $\{\ket{0},\ket{1}\}$ as settings $0$ and $1$ in $\mathcal{P}^n$, we get the violation \begin{align} v_w(n)=\frac{n-2}{n\times2^{n}}. \end{align} This algebraic violation, while works for all $\ket{W_n}$, is not the optimum violation. By optimizing over the four Euler angles in the two bases, we obtained close to optimal numerical violations of $\mathcal{P}^n$ (\ding{115} in Fig.~\ref{ghz_w_plot}) and $\mathcal{Q}^n_{n-1}$ (\ding{116} in Fig.~\ref{ghz_w_plot}) for W states. It can be seen from the plot that the violations of $\mathcal{P}^n$ is close to the upper bound derived from the geometrical measure of entanglement, $\frac{1}{2^{Eg(\ket{W_n})}}$. For GHZ states, we can follow the procedure given in section~\ref{bg} to find the bases. Note that the MPs of GHZ states with an even number of parties and an odd number of parties are different. For example, $\ket{+}$ is an MP of $\ket{GHZ_n}$ when $n$ is odd, but not when $n$ is even. Nevertheless, the MPs in both cases are all equally distributed along the equator of the Bloch sphere, allowing us to have a single expression for the bases as a function of $n$. The basis 1, which consists an MP and its antipodal point, is $\{\frac{1}{\sqrt{2}}(\ket{0}-e^{-\mathrm{i}\frac{\pi}{n}}\ket{1}),\frac{1}{\sqrt{2}}(\ket{1}+e^{\mathrm{i}\frac{\pi}{n}}\ket{0})\}$, and the basis 0 is $\{\frac{1}{\sqrt{2}}(\ket{0}-e^{\mathrm{i}\frac{(2n-1)\pi}{n(n-1)}}\ket{1}),\frac{1}{\sqrt{2}}(\ket{1}+e^{-\mathrm{i}\frac{(2n-1)\pi}{n(n-1)}}\ket{0})\}$. Calculating the violation as a function of $n$ (which is just the probability $P(0\ldots0|0\ldots0)$), we have (the \ding{110} line in Fig.~\ref{ghz_w_plot}) \begin{align} v_g(n)=\frac{1}{2^n}(1+cos(\frac{(2n-1)\pi}{n-1})). \end{align} This violation agrees with the best found by numerics. \begin{figure}[htbp] \begin{center} \includegraphics[width=230px,keepaspectratio=true]{Fig4.eps} \caption{Violations of $\mathcal{P}^n$ by the state $\ket{GHZ_n}$ (\ding{110}), with the numerical violations of $\mathcal{P}^n$ (\ding{115}) and $\mathcal{Q}^n_{n-1}$ (\ding{116}) of $\ket{W_n}$ as a function of $n$ (number of parties), comparing to $\frac{1}{2^{Eg(\ket{W_n})}}$ (\ding{108}) and $\frac{1}{2^{Eg(\ket{GHZ_n})}}$(\ding{117}).} \label{ghz_w_plot} \end{center} \end{figure} From Fig.~\ref{ghz_w_plot}, we can see that as $n$ increases, the violations are always well below $\frac{1}{2^{Eg}}$, which follows the trend we noticed in the earlier SDP examples. We also numerically optimized the value of $\mathcal{Q}^n_{n-1}$, which is always negative for GHZ states. This is in stark contrast to the situation for W states, where the violation of $\mathcal{Q}^n_{n-1}$ stays slightly below the violation of $\mathcal{P}^n$. One interpretation of this phenomenon is that $\mathcal{Q}^n_{d}$ is closely related to the degeneracy of the state, and can be used as a `witness' of degeneracy for these states. \section{Monogamy}\label{monogamy} \subsection{General Discussions} In sections~\ref{mobius} and~\ref{analytical}, we studied the nonlocality and entanglement for symmetric states from the perspective of different ``types'' in each context. There is another property, defined for both contexts, that highlights yet another interesting aspect of the relationship between nonlocality and entanglement: that is the concept of monogamy~\cite{PhysRevA.61.052306}~\cite{Toner08012009} (for a review see~\cite{springerlink:10.1007/s11128-009-0161-6}). As its name suggests, monogamy measures the ``exclusiveness'' of entanglement or correlations, that is, how well they can be shared. For example if two parties share a maximally entangled state or a maximally correlated Popescu-Rohrlich (PR) box~\cite{popescu1994quantum}, the entangled systems or PR box cannot be entangled or correlated to anything else. In recent years it has been recognized as a key ingredient to the usefulness of states for example in security and device independent security scenarios~\cite{PhysRevLett.97.170409,Pironio2010fk,1751-8121-44-9-095305,PhysRevLett.108.100401}. The idea being that if the correlations cannot be shared, that means that the eavesdropper is uncorrelated with the honest parties, so the information they share will not be leaked to the eavesdropper. Monogamy of entanglement is a property of a particular quantum state. It measures the intra-subgroup entanglement tradeoff with respect to a suitably chosen entanglement measure. The most famous such measure is the tangle $\tau$ introduced in~\cite{PhysRevA.61.052306}, which measures the entanglement across a bipartition. The CKW inequality, proposed in~\cite{PhysRevA.61.052306} as a conjecture and proved recently in~\cite{PhysRevLett.96.220503}, states that for all pure entangled states, the sum of all bipartite tangles between one party $A$ and $n$ parties $\{B_1,\ldots,B_n\}$ is less than or equal to the tangle between $A$ and all $B_i$ considered as a whole: \begin{align} \tau(\rho_{AB_1})+\tau(\rho_{AB_2})+\ldots+\tau(\rho_{AB_n})\leq\tau(\rho_{A(B_1\ldots B_n)}).\label{ckw} \end{align} Although it is known that symmetric states like the W state can saturate this inequality, not all states which saturate this inequality are symmetric. The monogamy of 3-qubit symmetric states have been studied recently~\cite{PhysRevA.85.012103}, using a different measure of quantum correlations, called the quantum deficiency (related to quantum discord~\cite{PhysRevLett.88.017901}). It was shown that SLOCC equivalent states do not necessarily have the same monogamy relation with respect to this measure. Here we focus on correlations of the measurement results directly (which we call simply ``monogamy of correlations"). Monogamy of correlations is normally defined in the context of correlations arising from probability distributions, without explicitly referring to quantum states and measurements. Intuitively, monogamy says that strong correlations cannot be shared. In a strict sense, we say an $n$-partite distribution, $P(a_1,\ldots,a_n|A_1,\ldots,A_n)$, is monogamous~\cite{PhysRevA.71.022101}~\cite{PhysRevLett.108.100401}, if the only nonsignaling extension to $n+1$ parties $P(a_1,\ldots,a_n,a_{n+1}|A_1,\ldots,A_n,A_{n+1})$ is the trivial one, i.e. such that \begin{align} &P(a_1,\ldots,a_n,a_{n+1}|A_1,\ldots,A_n,A_{n+1})\nonumber\\ =&P(a_1,\ldots,a_n|A_1,\ldots,A_n)P(a_{n+1}|A_{n+1}).\label{strict_mon} \end{align} For all possible measurement settings $A_k$ and $A'_k$ for party $k$, the nonsignaling condition can be stated as \begin{align} &P(a_1,\ldots,a_{k-1},a_{k+1}\ldots,a_{n}|A_1,\ldots,A_{k-1},A_{k+1},\ldots,A_{n})\nonumber\\ &=\sum_{a_k}P(a_1,\ldots,a_k,\ldots,a_{n}|A_1,\ldots,A_k,\ldots,A_{n})\nonumber\\ &=\sum_{a_k}P(a_1,\ldots,a_k,\ldots,a_{n}|A_1,\ldots,A'_k,\ldots,A_{n}).\label{no-signal} \end{align} That is, when tracing out one system, $k$, to get the marginal distributions, it does not matter which measurement setting $A_k$ is used. This strict sense of monogamy is guaranteed if an inequality reaches its algebraic maximum~\cite{PhysRevLett.97.170409}. Indeed, this fact is used to show monogamy for several states via several inequalities including GHZ states~\cite{PhysRevLett.97.170409}~\cite{PhysRevLett.108.100401}~\cite{Toner08012009}. However, the inequalities $\mathcal{P}^n$ and $\mathcal{Q}^n_d$ here cannot show strict monogamy in this way, simply because no quantum state can ever achieve the algebraic bound, as the bound is given by the entanglement. In the following subsection we will develop another set of inequalities for which this idea does work. Even if not demanding strict monogamy of correlations, it is possible to bound how well correlations can be shared. In~\cite{PhysRevLett.102.030403}, a bound is presented covering general nonsignaling theories by demanding tradeoffs of correlations in a multipartite setting, analogous to the monogamy of multipartite entanglement. To apply these results to our inequality, we will follow the prescription given in~\cite{PhysRevLett.102.030403}. First we rewrite our inequality to make all terms positive: \begin{align} \mathcal{P}^n&=P(0\ldots0|0\ldots0)\nonumber\\ &-(1-\sum_{a_1,\ldots,a_n\neq{0\ldots0}}P(a_1,\ldots,a_n|0\ldots1))\nonumber\\ &\vdots\nonumber\\ &-(1-\sum_{a_1,\ldots,a_n\neq{0\ldots0}}P(a_1,\ldots,a_n|1\ldots0))\nonumber\\ &-(1-\sum_{a_1,\ldots,a_n\neq{1\ldots1}}P(a_1,\ldots,a_n|1\ldots1)). \end{align} By keeping all the probabilities on the left hand side and moving everything else to the right hand side, we define the inequality \begin{align} \mathcal{P}^{n'}&=P(0\ldots0|0\ldots0)\nonumber\\ &+\sum_{a_1,\ldots,a_n\neq{0\ldots0}}P(a_1,\ldots,a_n|0\ldots1)\nonumber\\ &\vdots\nonumber\\ &+\sum_{a_1,\ldots,a_n\neq{0\ldots0}}P(a_1,\ldots,a_n|1\ldots0)\nonumber\\ &+\sum_{a_1,\ldots,a_n\neq{1\ldots1}}P(a_1,\ldots,a_n|1\ldots1)\nonumber\\ &\leq n+1. \end{align} Now we can partition the parties into two groups: group $A$ with $k$ parties and group $B$ with $n-k$ parties. Consider a single group $A$ which is possibly correlated with multiple identical $B^i$. The multiparty monogamy relation of~\cite{PhysRevLett.102.030403} tells us that for any nonsignalling probability distribution for $n>2$ \begin{align} \sum_{i=1}^{n-k+2}\mathcal{P}^{n'}(A,B^i)\leq (n-k+2) (n+1),\label{broad_mon} \end{align} where $i$ runs over the possible combinations of measurement settings of $n-k$ parties that make up each $B^i$. For $\mathcal{Q}_{d}^n$, we can treat the $d-1$ extra probabilities as marginals of probabilities involving $n$ parties: \begin{align} P(\underbrace{1\ldots1}_{n-d+1}|\underbrace{1\ldots1}_{n-d+1})=\sum_{b_1,\ldots,b_{d-1}}P(\underbrace{1\ldots1}_{n-d+1}\underbrace{b_1\ldots b_{d-1}}_{d-1}|\underbrace{1\ldots1}_{n}), \end{align} which leads to the inequality for $\mathcal{Q}_{d}^{n'}$: \begin{align} \mathcal{Q}_{d}^{n'}&=\mathcal{P}^{n'}+\sum_{a_1,\ldots,a_{n-1}\neq{1\ldots1}}P(a_1,\ldots,a_{n-1},b_1|\underbrace{1\ldots1}_{n})\nonumber\\ &\vdots\nonumber\\ +&\sum_{a_1,\ldots,a_{n-d+1}\neq{1\ldots1}}P(a_1,\ldots,a_{n-d+1},b_1,\ldots,b_{d-1}|\underbrace{1\ldots1}_{n})\nonumber\\ &\leq n+d. \end{align} Because the expression for $\mathcal{Q}_{d}^{n'}$ does not increase the number of settings for $B$, we have the monogamy inequality for $\mathcal{Q}_{d}^{n'}$ similar to (\ref{broad_mon}): \begin{align} \sum_{i=1}^{n-k+2}\mathcal{Q}_{d}^{n'}(A,B^i)\leq (n-k+2)(n+d).\label{broad_mon_q} \end{align} \subsection{New inequalities for Dicke states} We now introduce a set of inequalities which can show strict monogamy of Dicke states in the high $n$ limit. These are based on recent work by Heaney, Cabello, Santos and Vedral~\cite{1367-2630-13-5-053054} where they show that for the W state, it is possible to construct nonlocality tests and inequalities that are ``maximal'' in some sense, i.e. the violation of the inequality goes to the algebraic maximum in the $n\rightarrow\infty$ limit, thus mimicking perfect correlations of stabilizer states and the Mermin inequality~\cite{PhysRevA.77.062106}~\cite{PhysRevLett.65.1838}. The inequality introduced in~\cite{1367-2630-13-5-053054} by Heaney, Cabello, Santos and Vedral (hereafter referred to as the HCSV inequality), has the property that the larger $n$ is, the higher the violation becomes. Although the original HCSV inequality only works for W states, it can be extended as follows to cover all Dicke states. Following and extending the reasoning in~\cite{1367-2630-13-5-053054} for W state, if all $n$ parties measure in the $\sigma_z$ basis on a Dicke state $|S(n,k)\rangle$, $n-k$ of them will get result $0$ and the other $k$ will get result $1$ with certainty (though it is impossible to know who gets what). Now imagine that when $n-k-1$ parties get $0$ and the other $k-1$ parties get $1$, the remaining two decide instead to measure $\sigma_x$. In this case they will always get the same result. Since under LHV the results of one party should not depend on other parties' settings, this means that should any two chose to measure in $\sigma_x$, they would get the same result. If these results are given by an LHV distribution, this would mean that if all parties were to measure in $\sigma_x$ in the beginning, they should all get the same result. Since everything above occurs with certainty, we should always see, under LHV, that if all parties measure $\sigma_x$ they get the same result. However, simple calculation shows that this is not the case for all Dicke states. The associated Bell inequality is \begin{align} \mathcal{L}= &\sum P(\pi(\underbrace{0\dots0}_{n-k}\underbrace{1\dots1}_{k})|0\ldots0) \nonumber\\ -&\sum P(\pi(\underbrace{0\dots0}_{n-k-1}\underbrace{1\dots1}_{k-1}01)|\pi(\underbrace{0\dots0}_{n-2}11))\nonumber\\ -&P(0\ldots0|1\ldots1)-P(1\ldots1|1\ldots1) \leq 0,\label{heaneyext} \end{align} where the permutations in the second and third lines are over parties fixing the relationship between measurement settings and results, as with $\mathcal{P}^n$. To see that this cannot be violated under LHV it is sufficient to see that it cannot be violated for any deterministic strategy (i.e. taking marginal probabilities to be zero or one)~\cite{PhysRevA.64.032112}, since all LHV distributions can be considered as probabilistic mixtures of deterministic ones. It is not difficult to see that taking any one of the $P(\pi(\underbrace{0\dots0}_{n-k}\underbrace{1\dots1}_{k})|0\ldots0)$ to be one cannot be compatible with keeping all the negative terms zero. Since these are the only possible positive terms, and at most only one can be equal to one, for all deterministic local strategies the expression is non-positive and a violation is incompatible with LHV. For a Dicke state $\ket{S(n,k)}$, $\mathcal{L}$ is violated by $1-\frac{{n\choose k}}{2^{n-1}}$. As for the $W$ state considered in~\cite{1367-2630-13-5-053054}, this achieves the algebraic maximum in the limit of large $n$, imitating perfect correlations of GHZ and other stabilizer states. This also implies strict monogamy for the limit in $n$. We plot the violation of $\mathcal{L}$ for $\ket{S(n,\frac{n}{2})}$ and $|W_n\rangle$ in Fig.~\ref{heaney_plot}. We see that the W state reaches one more quickly, in keeping with its lower entanglement. \begin{figure}[htbp] \begin{center} \includegraphics[width=230px,keepaspectratio=true]{Fig5.pdf} \caption{A comparison of the violation of $\mathcal{L}$ (\ref{heaneyext}) for the states $\ket{S(n,\frac{n}{2})}$ (\ding{110}) and $\ket{W_n}$ (\ding{108}) as a function of $n$ (the number of parties).} \label{heaney_plot} \end{center} \end{figure} \section{Conclusions and Discussions} In this work we have studied the nonlocal properties of symmetric states as exposed by a set of inequalities. We have used the Majorana representation, numerics and semidefinite programming approaches to look at how classes of states can be identified using the inequalities, the scaling in $n$ for GHZ and W states and what we can say about monogamy of correlations that are seen. Concerning types of entangled states, we have been able to separate LUP and LU classes of states for four qubits using our inequalities, hence in a device independent way. The example states chosen also sit in different SLOCC classes. This was done by bounding the possible violation of inequalities using SDP techniques. Going above four qubits seems difficult as the numerics quickly get difficult with more parties, though simple basis checking numerics indicate that the W and GHZ states may be separated in this way. This furthers the discussion about how entanglement classifications can be interpreted using nonlocal features. On the one hand we have the general statement that degeneracy of MPs guarantees persistency of correlations \cite{PhysRevLett.108.210407} to subsystems. This is true for all states, not just specific examples such as those expended upon here. We see that certain ``example" states such as the $|000+\rangle$ and $|W\rangle$ states may be separated from less degenerate states using this fact. This can be compared to the robustness of nonlocality under system loss~\cite{PhysRevLett.108.110501}~\cite{PhysRevA.86.042113}. On the other hand, we also saw an example with the GHZ and T states where $\mathcal{Q}^n_d$ can be used to discriminate different classes, not related to degeneracy ($|T\rangle$ and $|GHZ\rangle$, both with degeneracy one). Intriguingly, we also remark that these states naturally appear in the phase space of spinor condensates~\cite{PhysRevLett.99.190408}, pointing to a potential interest of these ideas in many-body physics, for example to witness different phases of matter where standard order parameters fail. Existing connections between entanglement classes and symmetry could further be useful in this direction~\cite{PhysRevA.83.042332}. We also looked at what are the possible values violation can take. The first obvious statement with relation to entanglement was that the higher the entanglement is, the lower any possible violation of $\mathcal{P}^n$ and $\mathcal{Q}^n_d$ can be. At first this seems counterintuitive, but really it seems to stem from the simple fact that there is only one positive term - we later introduced larger inequalities with more positive terms based on the HCSV inequality~\cite{1367-2630-13-5-053054}, where the violation reaches its algebraic limit for all Dicke states in the high $n$ limit. We looked at how the violation of inequalities scale with $n$ for GHZ and W states. We see that W states fair much better for our inequalities, in contrast to the typical Mermin like inequalities where GHZ fairs better. We also look at the trends of the inequality violation with entanglement and see that this can be different. For W states and the $|000+\rangle$ state the violation increases with entanglement so that it gets closer to the upper bound ($\frac{1}{2^{E_g}}$), where as for GHZ states it goes down for higher $n$. We then looked at what can be said about the monogamy of the correlations exposed by our inequalities and chosen measurement settings. First, we see that $\mathcal{P}^n$ and $\mathcal{Q}^n_d$ are not suited to showing strict monogamy (that is, we cannot say violation at the level achieved by quantum states implies no correlations are shared with another party), since, by the fact that entanglement bounds the violation, any quantum violation cannot reach the algebraic limit. This may indicate that these inequalities are not so useful for device independent security for example, although bounds on correlation sharing less than these strict ones may be of interest. To this end, using techniques from~\cite{PhysRevLett.102.030403} we bound how much correlations can be shared with the inequalities. We then define new inequalities based on the HCSV inequality, where we see that all Dicke states are strictly monogamous in the limit of high $n$, as has been seen before for W states~\cite{1367-2630-13-5-053054}. In this sense the extreme nonlocality of GHZ and stabilizer states seems to be replicated by Dicke states in the large $n$ limit. It remains open how general this is for all symmetric states. One can also ask what other nonlocal properties can be inspected by inequalities $\mathcal{P}^n$ and $\mathcal{Q}^n_d$. Another property of multiparty correlations which is of interest, is whether it can be said to be ``genuine" or not - that is, whether the correlations at hand could be achieved by grouping the $n$ into subgroups or not. If not, we would say the correlations are genuinely $n$ party. The Svetlichny type inequalities~\cite{PhysRevD.35.3066} endeavor to identify this property - they should only be violated by genuinely $n$ party correlated states. Unfortunately it is not to hard to see that all the inequalities we use in this work do not have this property - it is possible to group parties together such that local states with respect to the new groupings can violate the inequalities. This can be easily seen by grouping the first $n-2$ parties and construct an LHV model by only using deterministic probabilities (probabilities equal to 0 or 1). The grouping makes it possible to set all negative terms to 0, and (one of) the positive term to 1. A stronger statement can be made by only grouping the first two terms - so that the weakest grouping still allows nonlocal correlations to violate all our inequalities. This is shown explicitly for $\mathcal{L}$ in the Appendix~\ref{proof}. In summary it seems that one must make a balanced choice over which inequalities will be useful depending on circumstances. We have seen that $\mathcal{P}^n$ and $\mathcal{Q}^n_d$ are interesting in terms of separating classes of states, and indeed it is known to be true that all entangled pure states will show some violation $\mathcal{P}^n$~\cite{PhysRevLett.109.120402}. However, their violation can never be high enough to make the strongest statements we would like about monogamy. They also do not say whether correlations are ``genuine" or not (even $\mathcal{L}$, with its many positive terms, does not show genuine nonlocality or be maximally violated for finite $n$). On the other hand inequalities based only on expectation values (which necessarily have many positive terms) can have maximal violation for any $n$, but they cannot see the nonlocality of all states - there are entangled states which do not violate any inequality based on expectation values, which do violate $\mathcal{P}^n$~\cite{PhysRevLett.88.210402} . In a similar situation to the role of different entanglement measures in entanglement theory, it seems unlikely that any single inequality will be able to capture all the nonlocal properties we might be interested in. \begin{acknowledgments} We thank Adel Sohbi for comments and discussions. We also thank Paul Jouguet for providing references regarding attacks on quantum key distribution systems. This work is supported by the joint ANR-NSERC grant ``Fundamental Research in Quantum Networks and Cryptography (FREQUENCY)''. \end{acknowledgments}
\section{Introduction}\label{sect-INTRO} A linear mapping between projectively embedded Euclidean spaces is called {\em central}, if its exceptional subspace is not at infinity. Such a linear mapping is in general not decomposable into a central projection followed by a similarity. Necessary and sufficient conditions for the existence of such a decomposition have been given in \cite{Have1} for arbitrary finite dimensions; cf. also \cite{Brau1}, \cite{Brau2}, \cite{Brau3}. However, those results do not seem to be immediately applicable on a {\em central axonometry}, i.e., a central linear mapping given via an axonometric figure. On the other hand, in a series of recent papers \cite{Pauk1}, \cite{Szab1}, \cite{Szab2} this problem of decomposition has been discussed for central axonometries of the Euclidean 3-space onto the Euclidean plane from an elementary point of vie \footnote{A lot of further references can be found in the quoted papers. . Loosely speaking, the concept of central axonometry is a geometric equivalent to the algebraic concept of a {\em coordinate matrix} for a linear mapping of the underlying vector spaces. However, from the results in \cite{Brau2} and \cite{Have1} it is also not immediate whether or not a given matrix describes (in terms of homogeneous Cartesian coordinates) a mapping that permits the above-mentioned factorization. The aim of this communication is to give a criterion for this. \vertsp Let $\bI$, $\bJ$ be finite-dimensional Euclidean vector spaces. Given a linear mapping $f : \bI \to \bJ$ denote by $f\ad:\bJ\to\bI$ its adjoint mapping. Then $f\ad \circ f$ is self-adjoint with eigenvalues \begin{displaymath} v_1 \ge \cdots \ge v_r > v_{r+1} = \cdots = v_n = 0. \end{displaymath} Here $r$ equals the rank of $f$ and $n=\dim \bI$. Moreover, each eigenvalue is written down repeatedly according to its multiplicit \footnote{For a self-adjoint mapping the algebraic and geometric multiplicities of an eigenvalue are identical. Hence we may unambiguously use the term `multiplicity'. . The positive real numbers $\sqrt{v_1},\ldots,\sqrt{v_r}$ are frequently called the {\em singular values} of $f$. The multiplicity of a singular value of $f$ is defined via the multiplicity of the corresponding eigenvalue of $f\ad \circ f$. It is immediate from the singular value decomposition that $f$ and $f\ad$ share the same singular values (counted with their multiplicities). See, e.g., \cite{Stra}. These results hold true, mutatis mutandis, when replacing $f$ by any real matrix, say $A$, and $f\ad$ by the transpose matrix $A^{\rm T}$. \section{Decompositions}\label{sect-KOO_FREI} When discussing central linear mappings it will be convenient to consider Euclidean spaces embedded in projective spaces. Thus let $\bV$ be an $(n+1)$-dimensional real vector space ($3\le n < \infty$) and $\bI$ one of its hyperplanes. Assume, furthermore, that $\bI$ is equipped with a positive definite inner product ($\cdot$) so that $\bI$ is a Euclidean vector space. In the projective space on $\bV$, denoted by $\Pro(\bV)$, we consider the projective hyperplane $\Pro(\bI)$ as the hyperplane at infinity. The absolute polarity in $\Pro(\bI)$ is determined by the inner product on $\bI$. Hence $\Pro(\bV) \setminus \Pro(\bI)$ is a projectively embedded Euclidean spac \footnote{We do not endow this space with a unit segment. . Similarly, let $\Pro(\bW) \setminus \Pro(\bJ)$ be an $m$-dimensional projectively embedded Euclidean space $(2\le m < n < \infty)$. Given a linear mapping \begin{equation}\label{f} f : \bV \to \bW \end{equation} of vector spaces then the associate (projective) linear mapping \begin{equation}\label{phi} \phi : \Pro(\bV) \setminus\Pro(\ker f)\to \Pro(\bW) \mbox{, } \RR\bx \mapsto \RR(f(\bx)) \end{equation} has $\Pro(\ker f)$ as its exceptional subspace. In the sequel we shall assume that \begin{equation}\label{cen-sur} \ker f \not\subset \bI \quad\mbox{and}\quad f(\bV) = \bW, \end{equation} or, in other words, that $\phi$ is central and surjectiv \footnote{This assumption of surjectivity is made `without loss of generality' in most papers on this subject. It will, however, be essential several times in this paper. . Obviously, (\ref{cen-sur}) is equivalent to \begin{equation}\label{combi} f(\bI) = \bW. \end{equation} We recall some results \cite{Brau2}, \cite{Have1}: If $\bT$ is any complementary subspace of $\ker f$ in $\bV$, then denote by \begin{equation} \psi_{\bT} : \Pro(\bV) \setminus\Pro(\ker f) \to \Pro(\bT) \end{equation} the projection with the exceptional subspace $\Pro(\ker f)$ onto $\Pro(\bT)$. The restricted mapping \begin{equation} \phi_{\bT} := \phi \mid \Pro(\bT) : \Pro(\bT) \to \Pro(\bW) \end{equation} is a collineation and \begin{equation} \phi = \phi_{\bT}\circ\psi_{\bT}; \end{equation} every decomposition of $\phi$ into a projection and a collineation is of this form. In the Euclidean vector space $\bI$ we have the distinguished subspace \begin{equation} \bE := f^{-1}(\bJ)\cap \bI. \end{equation} Write \begin{equation} f_\bE : \bE \to \bJ \mbox{, } \bx\mapsto f(\bx); \end{equation} this $f_\bE$ is well-defined and surjective, since $\bE\subset f^{-1}(\bJ)$ and $\ker f\not\subset \bE$. The subspace $\bT$ can be chosen with $\phi_\bT$ being a similarity if, and only if, the least singular value of $f_\bE$ has multiplicit \footnote{In \cite[Satz 10]{Brau2} this multiplicity is printed incorrectly as $2m-n-1$.} $\ge 2m-n+1$. Next, we assume that $\Pro(\bT)\not\subset \Pro(\bI)$ is orthogonal to $\Pro(\ker f)$. This means that $(\bT\cap\bI)^\bot \subset \ker f\cap\bI$ or $(\bT\cap\bI)^\bot \supset \ker f\cap\bI$. Hence $\psi_\bT$ is an {\em orthogonal central projection \footnote{The central projections used in elementary descriptive geometry are trivial examples of orthogonal central projections. . It is easily seen from \cite{Brau2} that $\phi$ permits a decomposition into an orthogonal central projection followed by a similarity if, and only if, all singular values of $f_\bE$ are equal. Finally, we are going to show that the crucial properties of $f_\bE$ can be read off from another mapping: Denote by \begin{equation} p : \bI \to \bE \end{equation} the orthogonal projection with the kernel $\bE^\bot\subset\bI$. Then \begin{equation}\label{f-p} (f_\bE \circ p) \circ (f_\bE \circ p)\ad = f_\bE \circ p \circ p\ad \circ (f_\bE)\ad = f_\bE \circ (f_\bE)\ad, \end{equation} since $p\ad$ is the natural embedding $\bE \to \bI$. Thus, by (\ref{f-p}) and the results stated in Section \ref{sect-INTRO}, $f_\bE$ and $(f_\bE \circ p)\ad$ have the same singular values (counted with their multiplicities). Hence, by the surjectivity of $f_\bE$ and (\ref{f-p}), all singular values of $f_\bE$ are equal if, and only if, there exists a real number $v>0$ such that \begin{equation}\label{ortho} (f_\bE \circ p) \circ (f_\bE \circ p)\ad = v\,\mbox{id}_\bJ. \end{equation} We shall use this in the next section. \section{A matrix characterization}\label{sect-KOORD} Introducing homogeneous Cartesian coordinates in $\Pro(\bV)$ is equivalent to choosing a basis $\{\bb_0,\ldots,\bb_n\}$ of $\bV$ such that $\{\bb_1,\ldots,\bb_n\}\subset\bI$ is an orthonormal system. The origin is given by $\RR\bb_0$ and the unit points are $\RR(\bb_0+\bb_1), \ldots, \RR(\bb_0+\bb_n)$. In the same manner we are introducing homogeneous Cartesian coordinates in $\Pro(\bW)$ via a basis $\{\bc_0,\ldots,\bc_m\}$. \begin{theo} Suppose that $f:\bV\to\bW$ is inducing a surjective central linear mapping $\phi$ according to formula (\ref{phi}). Let \begin{equation} A = \left( \begin{array}{lcl} a_{00} & \cdots & a_{0n} \\ {\rm\vdots} & {} & {\rm\vdots} \\ a_{m0} & \cdots & a_{mn} \end{array} \right) \end{equation} be the coordinate matrix of $f$ with respect to bases of $\bV$ and $\bW$ that are yielding homogeneous Cartesian coordinates. Write \begin{equation}\label{a_i} \ba_i := (a_{i1},\ldots,a_{in})\in \RR^n \mbox{ for all } i=0, \ldots, m \end{equation} and \begin{equation} \widetilde{A} := \left( \begin{array}{c} \ba_1- \frac{\ba_0\cdot\ba_1}{\ba_0\cdot\ba_0}\ba_0 \\ {\rm\vdots}\\ \ba_m - \frac{\ba_0\cdot\ba_m}{\ba_0\cdot\ba_0}\ba_0 \end{array} \right). \end{equation} Then the following assertions hold true: \begin{enumerate} \item $\phi$ is decomposable into a central projection followed by a similarity if, and only if, the least singular value of the matrix $\widetilde{A}$ has multiplicity $\ge 2m-n+1$. \item $\phi$ is decomposable into an orthogonal central projection followed by a similarity if, and only if, there exists a real number $v>0$ such that \begin{equation} \widetilde{A} \widetilde{A}^{\rm T} = \mbox{\rm diag}\,(v,\ldots,v). \end{equation} \end{enumerate} \end{theo} {\em Proof}. We read off from the top row of $A$ that \begin{displaymath} a_{00}x_0 + \cdots + a_{0n}x_n = 0 \end{displaymath} is an equation of $f^{-1}(\bJ)\neq \bI$ so that $\ba_0\cdot\ba_0\neq 0$. Write $\widetilde{f}:\bI\to\bJ$ for the linear mapping whose coordinate matrix with respect to $\{\bb_1,\ldots,\bb_n\}$ and $\{\bc_1,\ldots,\bc_m\}$ equals $\widetilde{A}$. A straightforward calculation shows that \begin{displaymath} \widetilde{f}(\bx) = f(\bx) \mbox{ for all } \bx\in \bE \end{displaymath} and \begin{displaymath} \widetilde{f}(a_{01}\bb_1+\cdots +a_{0n}\bb_n) = 0, \end{displaymath} i.e., $\bE^{\bot}\subset\ker\widetilde{f}$. Thus $\widetilde{f}$ equals the mapping $f_\bE \circ p$ discussed above. Now the proof is completed by translating formulae (\ref{f-p}) and (\ref{ortho}) into the language of matrices.\beweisende \vertsp We remark that (\ref{cen-sur}) and the linear independence of $\ba_1,\ldots,\ba_m$ are equivalent conditions. In contrast to the results in \cite{Pauk1}, \cite{Szab1}, \cite{Szab2}, the $\phi$-image of the origin $\RR\bb_0$ does not appear in our characterization. On the other hand, we have \begin{displaymath} f(\bE^\bot) = \RR((\ba_0\cdot\ba_0)\bc_0 +\cdots +(\ba_0\cdot\ba_m)\bc_m). \end{displaymath} In projective terms this 1-dimensional subspace of $\bW$ gives the {\em principal point} of the mapping $\phi$. Exactly if the principal point of $\phi$ equals the origin $\RR\bc_0$, then $\widetilde{A}$ arises from $A$ merely by deleting the top row and the leading column.
\section{Introduction} Let $\varphi(n)$ denote the Euler totient function of $n$. Lehmer \cite{Lehmer} asked whether there exist composite positive integers $n$ such that $\varphi(n)|n-1$. Integers which satisfy this \lq\lq Lehmer Condition" are sometimes referred to as Lehmer numbers, however no examples are known. Cohen and Hagis \cite{Cohen} have shown that any Lehmer numbers would necessarily have at least 14 prime factors, and computations by Pinch \cite{Pinch_Lehmer} show that any examples must be greater than $10^{30}$. Further, Luca and Pomerance \cite{Luca} have shown that if $\mathcal{L}(x)$ is the number of Lehmer numbers up to $x$ then, as $x \to \infty$, \[\mathcal{L}(x) \leq \frac{x^{1/2}}{(\log x)^{1/2+o(1)}}.\] Carmichael numbers are the composite integers $n$ which satisfy the congruence $a^n \equiv a \pmod{n}$ for every integer $a$. (Fermat's little theorem guarantees that any prime number $n$ satisfies this congruence.) Carmichael numbers were first characterized by Korselt \cite{Korselt} in 1899: \begin{korselt}A composite number $n$ is a Carmichael number if and only if $n$ is square-free, and for each prime $p$ which divides $n$, $p - 1$ divides $n - 1$. \end{korselt} Korselt did not find any Carmichael numbers, however. The smallest, 561, was found by Carmichael in 1910 \cite{Carmichael}. Carmichael also gave a new characterization of these numbers as those composite $n$ which satisfy $\lambda(n)|n-1$, where $\lambda(n)$, the Carmichael lambda function, denotes the size of the largest cyclic subgroup of $(\mathbb{Z}/n\mathbb{Z})^\times$. Since $\lambda(n)|\varphi(n)$ for every integer $n$, the Carmichael property can be viewed as a weakening of the Lehmer property. Every Lehmer number would also be a Carmichael number. In contrast to the Lehmer numbers, it is known, due to Alford, Granville and Pomerance \cite{Infinite}, that there are infinitely many Carmichael numbers. Pomerance \cite{Pomerance} also proves an upper bound for the number $C(x)$ of Carmichael numbers up to $x$, namely as $x \to \infty$, \begin{equation}C(x) \leq x^{1-\{1+o(1)\}\log\log\log x /\log\log x}, \label{car} \end{equation} and presents a heuristic argument that this is the true size of $C(x)$. Grau and Oller-Marc{\'e}n \cite{Grau} present other possible weakenings of the Lehmer property: looking at the sets of those $n$ such that $\varphi(n)|(n-1)^k$ for a fixed value of $k$ as well as the set of those $n$ for which $\varphi(n)|(n-1)^k$ for some $k$, that is all of the primes dividing $\varphi(n)$ also divide $n-1$. Note that this last set is a weakening of both the Lehmer and Carmichael properties, since $\lambda(n)$ and $\varphi(n)$ have the same prime divisors. Our results resolve several conjectures that Grau and Oller-Marc{\'e}n made in their paper. We focus primarily on this final set. Let $\kappa(n) = \rad{\varphi(n)}$ denote the product of the primes which divide the value $\varphi(n)$. (Note that $\kappa(n) = \rad{\varphi(n)} = \rad{\lambda(n)}$.) Let $\mathbb{K}(x) $ be the set of composite numbers $n \leq x$ which satisfy $\kappa(n)| n-1$, and let $K(x) = |\mathbb{K}(x)|$. (Observe that every prime number $p$ trivially satisfies $\kappa(p)|p-1$.) We prove that the upper bound \eqref{car} for $C(x)$ also applies for $K(x)$. We also present upper bounds for the number of $n \in \mathbb{K}(x)$ which are the product of a fixed number of primes, as well as several related conjectures and computations. \section{The Upper Bound} The condition for $n$ to be a member of $\mathbb{K}(x)$ is substantially weaker than that required for $n$ to be a Carmichael number, and computations (see Section \ref{sec:computations}) show that $K(x)$ appears to be substantially greater than $C(x)$. It is therefore somewhat surprising to find that these two functions have the same rough upper bound. Our proof of this fact is similar to the one for $C(x)$ in \cite{Pomerance}. \begin{theorem} Define $L(x) = \exp(\log x \frac{\log\log\log x}{\log\log x})$. Then as $x \rightarrow \infty$, \label{Main} \[K(x) \leq \frac{x}{L(x)^{1 + o(1)}}.\] \begin{proof} We consider first those integers $n \leq x$ which have a large prime divisor. Specifically, let $P(n)$ denote the largest prime divisor of $n$, and write $n = mp$ where $p = P(n)$. We restrict our attention to those $n$ with $P(n) > L(x)^2$, and let $K'(x) = \#\{n \in \mathbb{K}(x) \mid P(n) > L(x)^2\}$. If $n=mp$ is to satisfy $\kappa(n) | n-1$, then we must have $m \leq \frac{x}{p} $, and $m$ must be congruent to 1 $\pmod{ \rad{p-1}}$. Thus, for any fixed $p$ there are at most $1 + \lfloor\frac{x}{p\cdot\rad{p-1}}\rfloor$ possibilities for $m$. Requiring $n$ to be composite (thus $m \neq 1$) leaves us with at most $\frac{x}{p\cdot\rad{p-1}}$ possibilities. Thus we see that \begin{align} K'(x) &= \sum_{\substack{n = mp\leq x \\ p > L(x)^2 \\ \kappa(n)|n-1} } 1 \leq \sum_{\substack{p > L(x)^2} } \frac{x}{p \hspace{1mm} \rad{p-1} } \notag \\ &\leq \sum_{\substack{p > L(x)^2} } \frac{x}{(p-1) \hspace{1mm} \rad{p-1}}. \label{squarefree} \end{align} Now, we observe that for each prime $p$, the denominator in \eqref{squarefree} is a squarefull number, and that any squarefull number can be represented uniquely as $d \hspace{1mm} \rad{d}$ for some integer $d$. We can therefore replace this sum with a sum over all squarefull numbers: \begin{align*} \sum_{\substack{p > L(x)^2} } \frac{x}{(p-1) \hspace{1mm} \rad{p-1} } & \leq \sum_{\substack{d > L(x)^2 \\ d \text{ squareful}} } \frac{x}{d} . \end{align*} Using partial summation and the fact that \[\sum_{\substack{n \leq x \\ n \text{ squareful}} } 1 = \frac{\zeta(3/2)}{\zeta(3)} x^{1/2} + O(x^{1/3}),\] we see that \[K'(x) \leq \sum_{\substack{d > L(x)^2 \\ d \text{ squareful}} } \frac{x}{d} \ll \frac{x}{L(x)}.\] We may assume that $n>\frac{x}{L(x)}$, so to prove the theorem, it suffices to count those $n$ with $\frac{x}{L(x)} < n \leq x$ and $P(n) \leq L(x)^2$. We denote this count by $K''(x)$. Observe that every such $n$ has a divisor $d$ satisfying \begin{equation}\label{div} \frac{x}{L(X)^3} < d \leq \frac{x}{L(x)}.\end{equation} Write $n =md$, so $m\leq \frac{x}{d}$. Now, if $n = md$ is to satisfy $\kappa(md)|md-1$, we have $m \equiv 1 \pmod{\kappa(d)}$, and since $(n,\kappa(n)) = 1$ and $\kappa(d)|\kappa(n)$ we know $(d,\kappa(d))=1$. Thus the Chinese remainder theorem implies that there are at most $1 + \lfloor\frac{x}{d\kappa(d)}\rfloor$ possibilities for $m$. Thus \[K''(x) \leq \sideset{}{'}\sum \left( 1 + \frac{x}{d\kappa(d)} \right) \leq \frac{x}{L(x)} + \sideset{}{'}\sum \left\lfloor \frac{x}{d\kappa(d)} \right\rfloor ,\] where $\sideset{}{'}\sum$ denotes a sum over $d$ satisfying \eqref{div}. If $d\kappa(d) \leq x$ and $d$ satisfies \eqref{div}, then $\kappa(d) < L(x)^3$, so that \begin{align}K''(x) &\leq \frac{x}{L(x)} + \sideset{}{'}\sum \left\lfloor \frac{x}{d\kappa(d)} \right\rfloor \notag \\ &\leq \frac{x}{L(x)} + x \sum_{c \leq L(x)^3} \frac{1}{c} \sideset{}{'}\sum_{\kappa(d) = c}\frac{1}{d}. \label{doublesum} \end{align} We treat the inner sum in \eqref{doublesum} by partial summation: \begin{align} \sideset{}{'}\sum_{\kappa(d) = c}\frac{1}{d} = \frac{L(x)}{x} \sideset{}{'}\sum_{\kappa(d) = c} 1 + \int_{\frac{x}{L(x)^3}}^{\frac{x}{L(x)}}\hspace{2mm}\frac{1}{t^2} \hspace{2mm} \sideset{}{'}\sum_{\substack{\kappa(d) = c \\ d< t}} 1\hspace{2mm} dt. \label{parsum} \end{align} We are thus interested in obtaining an upper bound for $\mathcal{K}(t,c)$, the number of $d \leq t$ with $\kappa(d) = c$. \begin{lemma} \label{uniform} As $t \rightarrow \infty$, $\mathcal{K}(t,c) \leq \frac{t}{L(t)^{1+o(1)}}$ uniformly for all $c$. \end{lemma} Before proving the lemma, we see that using this upper bound in \eqref{parsum} gives us \begin{align*} \sideset{}{'}\sum_{\kappa(d) = c}\frac{1}{d} &\leq \frac{L(x)}{x} \mathcal{K}(\tfrac{x}{L(x)},c) + \int_{\frac{x}{L(x)^3}}^{\frac{x}{L(x)}}\hspace{2mm}\frac{1}{t^2} \hspace{2mm} \mathcal{K}(t,c)\hspace{2mm} dt \\ &\leq {L(\tfrac{x}{L(x)})^{-1+o(1)}} + \int_{\frac{x}{L(x)^3}}^{\frac{x}{L(x)}}\hspace{2mm}\frac{1}{tL(t)^{1+o(1)}} \hspace{2mm} dt \\ &= L(x)^{-1+o(1)} \end{align*} as $x \to \infty$. This can be used in \eqref{doublesum} to see that $K''(x) \leq \frac{x}{L(x)^{1+o(1)}}$. The theorem then follows immediately from our estimates of $K'(x)$ and $K''(x)$. It thus remains to prove Lemma \ref{uniform}. We may assume that $c \leq t$, otherwise $\mathcal{K}(t,c) = 0$. Then, for any $r > 0$ we can write: \begin{align*} \mathcal{K}(t,c) &= \sum_{\substack{d \leq t\\ \kappa(d) = c}} 1 \leq t^r \sum_{\substack{\kappa(d) = c}} d^{-r}\\ & \leq t^r \sum_{\substack{p|d \Rightarrow \rad{p-1}|c}} d^{-r} = t^r \prod_{\substack{\rad{p-1}|c}}\tfrac{1}{1-p^{-r}}. \end{align*} Assuming $r \geq 1/2 + \epsilon$ then \begin{align*} \prod_{\substack{\rad{p-1}|c}}\tfrac{1}{1-p^{-r}} &= \exp\left(\sum_{\substack{\rad{p-1}|c}}-\log({1-p^{-r})}\right) = \exp\left(\sum_{\substack{\rad{p-1}|c}}\hspace{3mm}\sum_{n=1}^\infty \frac{p^{-nr}}{n}\right) \\ &= \exp\left(\left(\sum_{\substack{\rad{p-1}|c}}p^{-r}\right) + O_\epsilon(1)\right). \end{align*} So we have \begin{align*} \mathcal{K}(t,c) &\ll_{\epsilon} t^r \exp \left(\sum_{\text{rad}(p-1) | c} p^{-r} \right) \leq t^r \exp \left(\sum_{\text{rad}(l) | c} l^{-r} \right) \\ & = t^r \exp \left(\prod_{p|c} (1-p^{-r})^{-1} \right) \leq t^r \exp\exp \left(\sum_{p|c}p^{-r} + O_\epsilon(1) \right) \end{align*} by applying this trick a second time. Now, $\sum_{p|c}p^{-r}$ is maximized when $c$ is the largest primorial up to $t$, in other words $c = p_1p_2 \cdots p_k < t$, where $p_i$ is the $i$th prime. Further, if $t$ is sufficiently large, then the prime number theorem implies that $p_k \leq 2\log(t)$ and thus \[ \sum_{p|c}p^{-r} \leq \sum_{p<2\log(t)}p^{-r} \] Choose $r = 1-(\log\log\log t)/(\log\log t)$. Thus for large $t$, we may choose $\epsilon = 1/4$. Then we have $t^r = \frac{t}{L(t)}$ and \[\sum_{p<2\log(t)}p^{-r} = O(\log\log t/\log\log\log t).\] Thus \begin{align*}\mathcal{K}(t,c) &\leq t^r \exp\exp \left(\sum_{p|c}p^{-r} + O_\epsilon(1) \right) \\ &= \frac{t}{L(t)}\exp\exp(O(\log\log t /\log\log\log t)) = \frac{t}{L(t)^{1 + o(1)}}, \end{align*} as $t \to \infty$, which completes the proof of the lemma. \end{proof} \end{theorem} \section{Bounds for integers in $\mathbb{K}(x)$ with $d$ prime factors} Since the integers satisfying our condition have a similar behavior to the Carmichael numbers assymptotically, it is natural to wonder if the behavior of those numbers with a fixed number of prime factors behaves similarly as well. Granville and Pomerance \cite{Granville} conjecture that the number, $C_d(x)$, of Carmichael numbers with exactly $d$ prime factors is $x^{1/d + o(1)}$ when $d\geq 3$, and as $x \to \infty$. This has not been proven for any $k$. However, Heath-Brown \cite{Heath} has shown that $C_3(x) \ll_\epsilon x^{7/20+\epsilon}$. Note that there are no Carmichael numbers with 2 prime factors. Let $K_d(x) = \#\{n \in \mathbb{K}(x), \omega(n)=d\}$ count the integers satisfying our condition up to $x$ with exactly $d$ prime factors. Using the same method as the first part of Theorem \ref{Main} we can prove \begin{theorem} Uniformly for $d\geq 2$ we have the bound $K_d(x) \ll x^{1-\frac{1}{2d}}$. \label{kd} \end{theorem} \begin{proof} Consider first those $n > x/2$. Since $n$ has $d$ prime factors, the largest prime factor must then satisfy $P(n) > (x/2)^{1/d}$. Applying the same argument used for integers $n$ with a large prime factor in Theorem \ref{Main}, we find that the total contribution of such integers is at most $O(x^{1-\frac{1}{2d}})$. Hence, $K_d(x) - K_d(x/2) \ll x^{1-\frac{1}{2d}}$. Now summing dyadically we have \[K_d(x) = \sum_{i=0}^\infty K_d(2^{-i}x)-K_d(2^{-i-1}x) \ll \sum_{i \geq 0} \left(\frac{x}{2^i}\right)^{1-\frac{1}{2d}} \ll x^{1-\frac{1}{2d}}. \] \end{proof} In contrast to the situation for Carmichael numbers, there do exist numbers satisfying our condition with two prime factors, and we can prove a substantially better bound than that of Theorem \ref{kd} in this case. As a matter of fact, their behavior appears to be like that conjectured for Carmichael numbers with a given number of prime factors. \begin{theorem} The numbers in $\mathbb{K}(x)$ with exactly two prime factors satisfy the bound $K_2(x) \leq x^{1/2}\exp\left(\frac{2(2\log x)^{1/2}}{\log\log x}\left(1+O\left(\tfrac{1}{\log\log x}\right)\right)\right)$. \end{theorem} \begin{proof}Write $n = pq \leq x$. Since $\kappa(pq) = \rad{(p-1)(q-1)}$ and $pq-1 = (p-1)(q-1)+(p-1)+(q-1)$ we have that $\kappa(pq)|pq-1$ if and only if $\rad{p-1}=\rad{q-1}$. Thus \begin{align*} K_2(x) &= \sum_{\substack{pq\leq x\\ \kappa(pq)|pq-1}} 1 \hspace{2mm}= \sum_{\substack{pq\leq x\\ \rad{p-1}=\rad{q-1}}} \hspace{-5mm} 1 \hspace{4mm} \leq \sum_{\substack{(m+1)(n+1) \leq x\\ \rad{m}=\rad{n}}} 1 \\ &\leq \sum_{\substack{mn \leq x\\ \rad{m}=\rad{n}}} 1 \hspace{2mm} \leq x^r \sum_{\substack{mn\leq x\\ \rad{m}=\rad{n}}} \frac{1}{(mn)^r} \end{align*} for any $r \geq 0$. We can rewrite this as a double sum: \begin{align*} x^r \hspace{-7mm} \sum_{\substack{mn\leq x\\ \rad{m}=\rad{n}}} \frac{1}{(mn)^r} &=x^r \sum_{m\leq x} \frac{1}{m^r}\sum_{\substack{n\leq x/m\\ p|m \text{ iff } p|n}} \frac{1}{n^r} \leq x^r \sum_{m\leq x} \frac{1}{m^r}\prod_{\substack{p|m}} \frac{\frac{1}{p^r}}{1-\frac{1}{p^r}} \\ &= x^r \sum_{m\leq x} \frac{1}{m^r\rad{m}^r}\prod_{\substack{p|m}} \frac{1}{1-p^{-r}}\\ &= x^r \sum_{m\leq x} \frac{1}{m^r\rad{m}^r}\exp\left(\sum_{\substack{p|m}}-\log\left(1-p^{-r}\right)\right)\\ & = x^r\sum_{m\leq x} \frac{1}{m^r\rad{m}^r}\exp\left(\sum_{\substack{p|m}} \sum_{j=1}^\infty \frac{p^{-jr}}{j}\right). \end{align*} As in the proof of Lemma \ref{uniform}, we can replace the condition $p|m$ above with $p\leq 2 \log x$, and $m\hspace{1mm}\rad{m}$ by a squareful integer $d$. We also set $r = 1/2$. Thus: \begin{align*} x^{1/2}\sum_{m\leq x} &\frac{1}{m^{1/2}\rad{m}^{1/2}}\exp\left(\sum_{\substack{p|m}} \sum_{j=1}^\infty \frac{p^{-j/2}}{j}\right) \\& \leq x^{1/2}\exp\left(\sum_{\substack{p\leq 2\log x}}\left( p^{-1/2} + \frac{1}{2p} + \sum_{j=3}^\infty \frac{p^{-j/2}}{j}\right)\right)\sum_{\substack{d\leq x^2 \\ d \text{ squarefull}}} \frac{1}{d^{1/2}}. \end{align*} By the prime number theorem we have \[\sum_{\substack{p\leq 2\log x}} p^{-1/2} = \text{li}\left(\left(2\log x\right)^{1/2}\right)\left(1+O\left(\frac{1}{\log\log x}\right)\right).\] So we can rewrite the expression above as \begin{align*} x^{1/2}\exp & \left( \text{li} \left((2\log x)^{1/2}\right)\left(1+O\left(\tfrac{1}{\log\log x}\right)\right ) + \tfrac{1}{2}\log\log\log x + O(1) \right) \hspace{-4mm}\sum_{\substack{d\leq x^2 \\ d \text{ squarefull}}}\hspace{-4mm}\frac{1}{d^{1/2}} \\ &= x^{1/2}\exp\left(\frac{2(2\log x)^{1/2}}{\log\log x}\left(1+O\left(\tfrac{1}{\log\log x}\right)\right)\right) \hspace{-4mm}\sum_{\substack{d\leq x^2 \\ d \text{ squarefull}}}\hspace{-4mm}\frac{1}{d^{1/2}}. \end{align*} By partial summation, we see that \[\sum_{\substack{d\leq x^2 \\ d \text{ squarefull}}}\hspace{-4mm}\frac{1}{d^{1/2}} = O(\log x),\] which can be absorbed into the existing error term in our equation, proving the theorem. \end{proof} Note that if we assume a strong form of the prime $k$-tuples conjecture, due to Hardy and Littlewood, we can show that this is fairly close to the actual size of $K_2(x)$. Their conjecture implies that the number of integers $m$ up to $x^{1/2}$ with both $m+1$ and $2m+1$ prime is asymptotically $cx^{1/2}/(\log x)^{2}$. Now, whenever both are prime, (and $m \neq 1$) we see that $\kappa((m+1)(2m+1)) = \rad{2m^2} = \rad{m}$, (since $m$ is necessarily even) and ${\rad{m}|(m+1)(2m+1)-1}$. Thus $K_2(x)$ would be at least of order $x^{1/2}/(\log x)^{2}$. \section{$k$-Lehmer Numbers} Grau and Oller-Marc{\'e}n \cite{Grau} define a $k$-Lehmer number to be an integer $n$ satisfying the condition $\varphi(n)|(n-1)^k$. (Note that they do not require $n$ to be composite, as we have in our definitions.) In their paper they make several conjectures about the counts of these $k$-Lehmer numbers. Our Theorem \ref{Main}, which shows in particular that $K(x) = O(\pi(x))$ (where $\pi(x)$ is the prime counting function) resolves four of these conjectures, Conjectures 8 (i)-(iv). Namely, this result proves Conjectures 8 (i),(ii) and (iv), while disproving (iii). Our methods, combined with the methods used in \cite{Lehmer23} to obtain a bound on the Lehmer numbers, can also be used to bound the counts of the $k$-Lehmer numbers. We let $\mathbb{L}_k(x)$ be the set of composite $n$ up to $x$ which satisfy ${\varphi(n)|(n-1)^k}$, and $L_k(x) = |\mathbb{L}_k(x)|$. (So Grau and Oller-Marc{\'e}n's function $C_k(x)=L_k(x)+\pi(x) +1$.) \begin{theorem} For $k \geq 2$ we have $L_k(x)\ll_k x^{1- \frac{1}{4k-1}}$. \end{theorem} \begin{proof} We consider three cases, based on the size of the largest prime divisor. We consider first those $n$, $x^{1-\frac{1}{4k-1}}< n \leq x$, which have $P(n)<x^{\frac{k}{4k-1}}$. Any such $n$ will have a divisor $d$ in the range $(x^\frac{k}{4k-1},x^\frac{2k}{4k-1})$. Write $n = md$, so $m \leq /d$ and since $\varphi(md)|(md-1)^k$, we see that $(md-1)^{k} \equiv 0 \pmod{\varphi(d)}$. Now, for any positive integer $N$, the number of residue classes $r \pmod{N}$ with $r^k \equiv 0 \pmod{N}$ is at most $N^\frac{k-1}{k}$. Thus, for any fixed $d$, using the fact that $(d,\varphi(d))=1$, we see that $m$ must be in one of at most $\varphi(d)^{\frac{k-1}{k}}$ residue classes mod $\varphi(d)$, giving us at most \[\varphi(d)^{\frac{k-1}{k}}\left\lceil\frac{x}{d\varphi(d)}\right\rceil \leq \varphi(d)^{\frac{k-1}{k}}\left(1+\frac{x}{d\varphi(d)}\right)\] choices for $m$. Summing over all $d$ in the range $I = (x^\frac{k}{4k-1},x^\frac{2k}{4k-1})$, we get \begin{align*} \sum_{d \in I} \varphi(d)^{\frac{k-1}{k}}\left(1+\frac{x}{d\varphi(d)}\right) &\leq \sum_{d \in I} d^\frac{k-1}{k} + \frac{x}{d^{1+\frac{1}{k}}}\left(\frac{d}{\varphi(d)}\right)^{\frac{1}{k}} \\ & \leq \sum_{d \in I} d^\frac{k-1}{k} + \sum_{d \in I} \frac{x}{d^{1+\frac{1}{k}}}\left(\frac{d}{\varphi(d)}\right). \\ \end{align*} The first sum is $\ll x^{1-\frac{1}{4k-1}}$. Now, using partial summation on the second sum and the fact that $\sum_{t \leq x} \frac{t}{\varphi(t)} = O(x)$, we get \begin{align*} \sum_{d \in I} \frac{x}{d^{1+\frac{1}{k}}}\left(\frac{d}{\varphi(d)}\right) &\ll \frac{x}{x^{(\frac{2k}{4k-1})({1+\frac{1}{k}})}}\sum_{d\leq x^\frac{2k}{4k-1}}\frac{d}{\varphi(d)}+ x\int_{x^\frac{k}{4k-1}}^{x^\frac{2k}{4k-1}}\frac{1}{t^{2+\frac{1}{k}}}\sum_{i<t}\frac{t}{\varphi(t)}dt\\ &\ll \frac{x}{x^{(\frac{2k}{4k-1})({1+\frac{1}{k}})}}\left(x^\frac{2k}{4k-1}\right) + x\int_{x^\frac{k}{4k-1}}^{x^\frac{2k}{4k-1}}\frac{1}{t^{1+\frac{1}{k}}}dt\\ &\ll_k x^{1-\frac{2}{4k-1}} + \frac{x}{x^{(\frac{k}{4k-1})({\frac{1}{k}})}} \ll x^{1-\frac{1}{4k-1}}. \end{align*} In the second case we consider those $n$ with $x^{\frac{k}{4k-1}}<P(n) \leq x^{\frac{2k}{4k-1}}$. In this case $n$ again has a divisor in the range $(x^\frac{k}{4k-1},x^\frac{2k}{4k-1})$, namely $p$, and the above argument applies verbatim. Finally we've reduced to the case that $P(n)>x^{\frac{2k}{4k-1}}$, and the argument used for large primes in our main theorem gives us that the number of $n$ with $\kappa(n)|n-1$ and $P(n)> x^{\frac{2k}{4k-1}}$ is at most $x^{1-\frac{k}{4k-1}}$, hence for those $n$ in $\mathbb{L}_k(x)$ as well, and our result follows. \end{proof} We note that it may be possible to improve upon this bound by using techniques developed in more recent papers to obtain better bounds on the Lehmer numbers. \section{Computations and Conjectures} \label{sec:computations} Table \ref{table} shows the values of $K(x)$ we computed for increasing powers of 10, compared with values of $C(x)$, computed by Richard Pinch \cite{Pinch}. Our computations were done using trial divison, in which a candidate number, $n$, was rejected as soon as soon as it was found to be nonsquarefree, or to have a prime divisor $p$, which failed to satisfy $\rad{p-1}|n-1$. \begin{table}[ht] \caption{Values of $C(x)$ and $K(x)$ to $10^{11}$.} \label{table} \begin{tabular}{|l|l|l|} \hline $n$ & $C(10^n)$ & $K(10^n)$ \\ \hline 2 & 0 & 4 \\ 3 & 1 & 19 \\ 4 & 7 & 103 \\ 5 & 16 & 422 \\ 6 & 43 & 1559 \\ 7 & 105 & 5645 \\ 8 & 255 & 19329 \\ 9 & 646 & 64040 \\ 10 & 1547 & 205355 \\ 11 & 3605 & 631949 \\ \hline \end{tabular} \end{table} Despite the similar asymptotic bounds that we have for $C(x)$ and $K(x)$, it is clear that $K(x)$ is growing substantially faster, which leads to the conjecture: \begin{conjecture} $\lim_{x \to \infty} K(x)/C(x) = \infty$. \end{conjecture} At the moment, however, we are unable to prove even the much weaker conjecture: \begin{conjecture} $\lim_{x \to \infty} K(x) - C(x) = \infty$. \end{conjecture} \section*{Acknowledgments} I would like to thank my advisor, Carl Pomerance, for suggesting the problem and for his invaluable guidance and encouragement throughout the development of this paper. \bibliographystyle{amsplain}
\section{Introduction} The area of communication complexity deals with the amount of communication required for solving computational problems with distributed input. In the two-party \emph{simultaneous message passing (SMP)} setting of communication complexity, two \emph{players} Alice and Bob receive inputs~$x$ and~$y$, respectively, and each sends a message to a third party, the \emph{referee}. Using those messages, the referee computes the output value. When the goal is to compute certain function~$f$, the success is measured by the probability that the output value of a \emph{communication protocol} equals $f(x,y)$. The \emph{cost} of a communication protocol is the total number of bits sent by the players to the referee. Shared randomness is a crucial resource in communication complexity. When Alice and Bob have it, they can use a mixed strategy in order to compute $f$; in particular, the minimax principle implies that the worst-case and the average-case complexities are equal in this case. It is known that without shared randomness the model becomes considerably weaker, and the gap between the worst-case and the average-case complexities of a communication problem can be arbitrary large (constant vs.~$\Omega(\sqrt{n})$ in the case of the equality function, as shown by Newman and Szegedy~\cite{NS96_Pub}). The SMP model of communication is the weakest among those that have been studied widely. Nevertheless, it is arguably the right model to look at when the goal is to investigate shared randomness. That is because whenever communication between the players is possible (which is the case for all other commonly studied models, but not for SMP), the first player can append~$O(\log n)$ bits of \emph{private} randomness to the first message that is sent to the others, and that would not affect the cost of the protocol significantly, as poly-logarithmic cost is usually viewed as efficient. Those random bits are now known to all the participants, and can be used in place of shared randomness. It is known due to Newman~\cite{N91_Pri} that~$O(\log n)$ bits of shared randomness are always enough; therefore, shared randomness does not make much difference in any model that allows direct communication between the players. In this paper we study shared randomness in the context of the \emph{multi-party version} of the SMP model, where the number of players~$k$ is three or larger (the referee is not counted as a player), the input has~$k$ fragments and each fragment is known to exactly one player-- this regime of distributing input between the players is usually called ``number in hand.'' This model can be viewed as a natural generalization of the two-party model. We demonstrate several interesting (and somewhat surprising) properties of shared randomness when the number of players is at least three, that have no direct analogues in the two-party case. \subsection{Previous work} In~\cite{Y03_On_th}, Yao generalized the technique of \emph{quantum fingerprints}~\cite{BCWW01_Qu} to show that every classical two-party SMP protocol that uses shared randomness and sends~$c$ bits can be simulated by a quantum protocol without shared randomness that sends~$2^{O(c)}\log n$ qubits. This naturally raised the question whether quantum communication can always replace shared randomness-- that is, whether any communication problem that can be solved by a classical SMP protocol of poly-logarithmic length using shared randomness can also be solved by a quantum protocol of poly-logarithmic length without shared randomness. The question was addressed by Gavinsky, Kempe, Regev and de~Wolf in~\cite{GKRW06_Bou}, where they demonstrated a two-party \emph{relational} communication problem that can be solved by a classical protocol of cost~$O(\log n)$ that uses shared randomness, but requires~$n^{\Omega(1)}$ qubits in order to be solved by a quantum protocol without shared randomness. In the same work a question was posed whether a similar separation is possible via a \emph{functional} problem. \section{Our results} Two main results of this work are the following. First, we establish a hierarchy of modes of shared randomness (Section~\ref{s_srand}). In the $k$-party SMP model, we consider \emph{$t$-shared randomness} for~$2\le t\le k$, where every set of~$t$ players shares a random string. The~$k$-shared randomness is the usual, unrestricted shared randomness and the strongest mode in the hierarchy, and a smaller value of~$t$ gives a weaker form of shared randomness. The~$(k-1)$-shared randomness could be also called ``randomness on the forehead.'' Below~$2$-shared randomness, we also consider an even more restricted mode of shared randomness which we call \emph{XOR-shared randomness}, where the~$k$ players receive uniformly random~$k$-tuples of bits whose parity is~$0$. The precise definitions of these modes of shared randomness will be given in Section~\ref{s_srand}. We will show that this is a proper hierarchy; i.e., we show exponential separations between its levels. One of the problems that we study in this context is the multi-party equality function, and we show (Claim~\ref{c_k-eq}) that it can be solved by a protocol of constant length that uses XOR-shared randomness. We believe that this result might be of independent interest, due to the importance of the equality function. Second, we demonstrate a promise function whose classical communication complexity is constant if the strongest form of shared randomness is available, but whose quantum communication complexity is~$n^{\Omega(1)}$ if no shared randomness is available (Section~\ref{s_vs}). Moreover, the quantum complexity remains~$n^{\Omega(1)}$ even if the protocol can use $(k-1)$-shared randomness (randomness on the forehead). Our second result is closely related to~\cite{GKRW06_Bou}: We demonstrate a \emph{promise function} that can be solved efficiently in the classical model with shared randomness, but not in the quantum model without it. This answers the main open problem posed in~\cite{GKRW06_Bou} for the case of three or more players. We note that the question remains wide open in the two-player case. Our second result is also related to the aforementioned work by Yao~\cite{Y03_On_th}, where it was shown, informally speaking, that shared randomness can be replaced by quantum communication with (at most) exponential overhead. In this work we demonstrate a (functional) communication problem that can be solved by a three-bit three-party classical protocol with shared randomness but requires~$n^{\Omega(1)}$ qubits without shared randomness (or even with randomness on the forehead). Accordingly, the possibility to simulate shared randomness by quantum communication is a unique feature of the two-party model; with more than two players, the possible advantage of shared randomness over quantum communication is not bounded by any function. \subsection{Technical statements} In the first part, we prove the following. \newcommand{\theohie} Let $k\ge3$. Then, \begin{itemize} \item For each $t\in\{3,\dots,k\}$, there exists a~$k$-party promise function that can be solved by a protocol of cost~$t$ in the SMP model with classical communication and~$t$-shared randomness but requires~$\Omega(tn^{1/t})$ qubits in the SMP model with quantum communication and $(t-1)$-shared randomness. \item There exists a~$k$-party total function that can be solved by a protocol of constant cost in the classical SMP model with $2$-shared randomness but requires~$\Omega(\sqrt{n})$ bits of communication in the classical SMP model with XOR-shared randomness. \item There exists a~$k$-party total function that can be solved by a protocol of constant cost in the classical SMP model with XOR-shared randomness but requires~$\Omega(\sqrt{n})$ bits of communication in the classical SMP model without shared randomness. \end{itemize} } \begin{theorem} \label{t_hie} \theohie \end{theorem} In the second part of this work, we study the following natural communication problem. For a bit string~$x$, we denote by~$\abs{x}$ its Hamming weight, i.e.\ the number of 1s in~$x$. \begin{definition}[Gap-Parity] Let~$x_1,\dots,x_k$ be~$n$-bit strings such that~$\abs{x_1\oplus\dots\oplus x_k}\notin[n/3,2n/3]$. Then we define~$\GPk(x_1,\dots,x_k)=0$ if~$\abs{x_1\oplus\dots\oplus x_k}<n/2$ and~$\GPk(x_1,\dots,x_k)=1$ otherwise. \end{definition} Note that in the SMP model with classical communication and shared randomness, $\GPk$ has a trivial solution, where each player sends only one bit to the referee. We will demonstrate that for~$k\ge3$, $\GPk$ cannot be solved efficiently by a quantum protocol without shared randomness. Moreover, we show that the Gap-Parity problem has no efficient solution with quantum communication even with randomness on the forehead (cf.\ Section~\ref{s_srand}). \newcommand{\theomain} Let~$k\ge3$. Using shared randomness, the~$k$-party promise function~$\GPk$ can be solved by a classical SMP protocol of cost~$k$ where each player sends a bit. For~$2\le t\le k-1$, in the SMP model with quantum communication with~$t$-shared randomness, the complexity of~$\GPk$ is~$\Omega(kn^{1-t/k})$. In particular, in the SMP model with quantum communication without shared randomness, the complexity of~$\GPk$ is~$\Omega(kn^{1-2/k})$. } \begin{theorem} \label{t_main} \theomain \end{theorem} \section{Preliminaries} \label{s_prel} For any~$n$-dimensional vector~$v$, we will write~$v(j)$ to denote its~$j$th coordinate, and for any~$S\subseteq[n]$ we will use~$v_S$ to denote the restriction of~$v$ to the coordinates that are elements of~$S$. We use~$\bar{0}$ or~$\bar{1}$ to denote the vectors of all $0$s or all $1$s, respectively, when its length is clear from the context. For any finite set~$W$, let~$\U(W)$ denote the uniform distribution over the elements of~$W$. \subsection{Quantum measurements} Unless stated otherwise, we will represent quantum states by their density matrices. We will write~$\E_i{\sigma_i}$ or even $\E{\sigma_i}$ to denote the mixed state~$(1/k)\sum_i\sigma_i$. Given a matrix~$M$, we denote by~$\norm{M}_1$ the \emph{trace norm} of~$M$, defined as the sum of the singular values of~$M$. It is known that given two quantum states~$\sigma_0$ and~$\sigma_1$, the optimal probability with which a quantum measurement can correctly distinguish between~$\sigma_0$ and~$\sigma_1$ equals $1/2+\norm{\sigma_0-\sigma_1}_1/4$. We will use the following special case of the ``random access code argument''~\cite[Lemma~2.2]{GKRW06_Bou}, which is a slight generalization of~\cite[Theorem~2.3]{N99_Op_Lo} (see also~\cite{ANTV02_De}). \begin{claim} \label{c_rac} Let~$X \sim \U(\zon)$. Suppose for each instantiation~$X=x$ we have a quantum state~$\rho_x$ of~$q$ qubits. Let~$\sigma_a^j$ be the expectation of~$\rho_X$ conditional upon~$X(j)=a$, for~$j\in[n]$ and~$a \in \zo$. Then~$\sum_{j=1}^n\norm{\sigma_0^j-\sigma_1^j}_1^2 \in O(q)$. \end{claim} \begin{proof} Let~$h(p)$ be the binary entropy function: $h(p)=-p\log_2 p-(1-p)\log_2(1-p)$. Lemma~2.2 of~\cite{GKRW06_Bou} implies that under the assumption of the claim, it holds that $\sum_{j=1}^n \bigl(1-h(1/2-\norm{\sigma_0^j-\sigma_1^j}_1/4)\bigr) \le q$. The claim follows because~$1-h(1/2-x/4)\ge x^2/(8\ln2)$ for~$0\le x\le2$. \end{proof} Now let us consider the situation where a quantum measurement is performed in order to predict the parity of several independent binary variables. \begin{claim} \label{c_xor} For every~$i\in[m]$, let~$\sigma_0^i$ and~$\sigma_1^i$ be quantum states of equal dimension. For~$a\in\zo$, let $\rho_a\deq\E_{\alpha_1\oplus\dots\oplus\alpha_m=a}{\sigma_{\alpha_1}^1\otimes\dots\otimes\sigma_{\alpha_m}^m}$. Then $\norm{\rho_0-\rho_1}_1 =(1/2^{m-1})\prod_{i=1}^m\norm{\sigma_0^i-\sigma_1^i}_1$. \end{claim} \begin{proof} Write: \[ \rho_0-\rho_1= \frac{1}{2^{m-1}}(\sigma_0^1-\sigma_1^1)\otimes\dots\otimes (\sigma_0^m-\sigma_1^m), \] and the claim follows from the fact that the trace norm is multiplicative with respect to the tensor product. \end{proof} \subsection{Communication complexity} In this work we are interested in the following model of communication complexity. \begin{definition}[Multi-party SMP] The~$k$-party simultaneous message passing (SMP) model involves~$k+1$ parties: $k$ players~$\A_1,\dots,\A_k$ and a referee. For every~$i\in[k]$, player~$\A_i$ gets input~$x_i$. They each send one message to the referee, who uses the content of all~$k$ messages to compute the output value. A communication protocol describes the action of each participant. The \emph{cost} or \emph{complexity} of a protocol is the total length of the messages sent by players~$A_1,\dots,A_k$ to the referee. We say that a protocol solves a computational problem defined over~$k$ input values if the referee gives a correct answer with probability at least~$2/3$ for each possible input. \end{definition} In this paper we will consider several further modifications of the SMP model: \begin{itemize} \item In the \emph{quantum} SMP model, the players~$\A_1,\dots,\A_k$ are allowed to send quantum messages, and the referee can perform any quantum measurement in order to determine the output value. \item In the SMP model \emph{with shared randomness}, the players~$\A_1,\dots,\A_k$ have free access to the same string of random bits that were chosen independently from the input values. \item In Section~\ref{s_srand}, we will define a \emph{hierarchy of modes of shared randomness} in multi-party protocols (where the strongest mode is the standard one, as described above). We demonstrate exponential separations between the levels of the hierarchy (i.e., the hierarchy is proper). \end{itemize} We call a communication protocol \emph{efficient} if its cost is poly-logarithmic in the length of input. \subsection{Read-$k$ families of functions} Let us consider the following model of dependence among random variables. \begin{definition}[Read-$k$ families] Let~$X_1,\dots,X_m$ be independent random variables. For~$j \in [r]$, let~$P_j\subseteq[m]$ and let~$f_j$ be a Boolean function of~$(X_i)_{i\in P_j}$. If every~$i\in [m]$ belongs to at most~$k$ among the~$r$ sets~$P_1,\dots,P_r$, then the random variables~$Y_j=f_j((X_i)_{i \in P_j})$ are called a \emph{read-$k$ family}. \end{definition} The following lemma is due to Finner~\cite{F92_A_Ge}. \begin{lemma}[Finner~\cite{F92_A_Ge}] \label{l_Hold} Let~$Y_1,\dots,Y_n$ be a read-$k$ family of random variables taking non-negative values. Then \[ \E*{\prod_{i=1}^nY_i}\le\prod_{i=1}^n\sqrt[k]{\E{Y_i^k}}. \] \end{lemma} Note that the generalized H\"older inequality implies that $\E{\prod Y_i}\le\prod\sqrt[n]{\E{Y_i^n}}$ in general, without making any independence assumption. This corresponds to choosing~$k=n$ in the lemma. On the other hand, when~$k=1$ (i.e., $Y_1,\dots,Y_n$ are mutually independent), the expectation of their product equals the product of their expectations. Accordingly, Lemma~\ref{l_Hold} gives a natural interpolation between these two extreme cases. \section{Hierarchy of shared randomness in multi-party protocols} \label{s_srand} When there are more than two players, it is possible to give the players access to shared randomness in several different ways. Most naturally, the parties may have free access to the same string of random bits-- we call this mode \emph{unrestricted shared randomness}. Note that this mode of shared randomness is often implicitly assumed to be available to the players; for example, unrestricted shared randomness is required in order to be able to use \emph{mixed strategies}, and therefore applicability of the minimax principle in multi-party communication depends on it. Let~$k\ge2$ be the number of players. For every~$t\in\set{2,\dots,k}$, we can define the mode of \emph{$t$-shared randomness}, where every~$t$ players share their own string of random bits. The $k$-shared randomness is the same thing as the unrestricted shared randomness. Sometimes we will refer to the~$(k-1)$-shared mode as \emph{randomness on the forehead}. We will also consider \emph{XOR-shared randomness}, where every player~$\A_i$ is given access to an arbitrarily long random string~$r_i$, such that every~$(k-1)$ strings~$r_i$ are uniform and mutually independent but the bitwise XOR of~$r_1,\dots,r_k$ is~$0$ everywhere. If we consider the case of~$k=2$, we can see that XOR-shared and $2$-shared modes are the same. For~$k=3$, we already have three modes of shared randomness: XOR-shared, $2$-shared and $3$-shared. We will see below that these three modes offer different computational power. In general, $t$-shared randomness is always at least as strong as~$(t-1)$-shared randomness, as the latter can always be emulated using the former. Also, XOR-shared randomness can be emulated in the $2$-shared mode; to do that, let $r_i$ equal the bit-wise XOR of the random string shared between~$\A_{i-1}$ and~$\A_i$ and the random string shared between~$\A_i$ and~$\A_{i+1}$, where the subscripts are interpreted modulo~$k$. Now the strings $r_1,\dots,r_k$ are distributed as required by the definition of XOR-shared mode. One interesting example that demonstrates usefulness of XOR-shared randomness when~$k\ge3$ is the \emph{multi-party equality function}; that is, the total Boolean function of~$k$ arguments~$x_1,\dots,x_k$ that takes value~$1$ if and only if~$x_1=x_2=\dots=x_k$. \begin{claim} \label{c_k-eq} For any~$c\in\NN$, there exists a classical protocol for the~$k$-party equality function, where XOR-shared randomness is used, each player sends~$c$ bits to the referee, and the following holds: \begin{itemize} \item If~$x_1=x_2=\dots=x_k$, then the referee's answer is always~$1$; \item otherwise, the referee's answer is~$0$ with probability~$1-1/2^c$. \end{itemize} \end{claim} \begin{proof} For~$r,x\in\zo^n$, let~$r\cdot x$ be the inner product of~$r$ and~$x$ in finite field~$\GF(2)$: $r\cdot x\deq\bigoplus_{i:r(i)=1}x(i)$. Consider the following protocol~$\Pk$, where the players use XOR-shared randomness and each of them sends a single bit to the referee: \begin{enumerate}[1.] \item For all~$i\in[k]$, the~$i$th player uses his random string~$r_i\in\zo^n$ and computes~$m_i\deq r_i\cdot x_i$, then sends~$m_i$ to the referee. \item The referee outputs~$\neg(m_1\oplus\dots\oplus m_k)$. \end{enumerate} By the definition of XOR-shared randomness, we have that~$r_1\oplus\dots\oplus r_k=\bar0$. Therefore~$(r_1\oplus\dots\oplus r_k)\cdot x_k=0$, and we can write \begin{align} & m_1\oplus\dots\oplus m_k \nonumber \\ &=r_1\cdot x_1\oplus\dots\oplus r_k\cdot x_k \nonumber \\ &=r_1\cdot x_1\oplus\dots\oplus r_k\cdot x_k \oplus(r_1\oplus\dots\oplus r_k)\cdot x_k \nonumber \\ &=r_1\cdot(x_1\oplus x_k)\oplus\dots\oplus r_{k-1}\cdot(x_{k-1}\oplus x_k). \label{m_keq} \end{align} Note that~$r_1,\dots,r_{k-1}$ is a uniformly random~$(k-1)$-tuple of~$n$-bit strings, and therefore the rightmost part of~(\ref{m_keq}) equals~$1$ with probability exactly~$1/2$ if at least one of the~$k-1$ values~$x_1\oplus x_k,\dots,x_{k-1}\oplus x_k$ is different from~$\bar0$. If, on the other hand, $x_1=x_2=\dots=x_k$ then(\ref{m_keq}) equals~$0$ with certainty. Accordingly, $\Pk$ outputs ``$1$'' whenever $x_1=x_2=\dots=x_k$, and otherwise it outputs ``$0$'' with probability~$1/2$. To get a protocol as promised by our claim, we can run~$c$ independent instances of~$\Pk$ in parallel and output ``$1$'' if and only if all~$c$ instances answered ``$1$''. \end{proof} We are now prepared to prove that the modes of shared randomness form a proper hierarchy when~$k\ge3$. \begin{reptheorem}{t_hie} \theohie \end{reptheorem} \begin{proof} To prove the first part of the theorem, consider the~$t$-party problem~$\GP_t$, letting the players~$\A_1,\dots,\A_t$ receive the corresponding fragments of input. (The other~$(k-t)$ players do not receive any input.) Theorem~\ref{t_main}, which will be proved in the next section, implies the result in this case. The second part follows from considering the two-party equality problem, when the input is distributed between~$\A_1$ and~$\A_2$ and the other players do not receive any input. It is clear that this problem can be solved with constant cost in the classical SMP model with $2$-shared randomness. Now suppose that there exists a protocol of cost~$c$ in the classical SMP model with XOR-shared randomness, and we will prove that~$c=\Omega(\sqrt{n})$. Note that in this model, the random strings given to~$\A_1$ and~$\A_2$ are uniform and independent, although they are correlated with the random strings given to the other players. Such a protocol can be transformed without changing the cost to a protocol in the two-party classical SMP model where the two players do not share randomness but each player shares randomness with the referee, because in the latter model, the referee can generate all the messages which would have been generated by players~$\A_3,\dots,\A_k$. By the same technique used by Newman~\cite{N91_Pri}, this protocol can be further transformed to a protocol of cost~$O(c+\log n)$ in the two-party classical SMP model without shared randomness at all. It was shown by Newman and Szegedy~\cite{NS96_Pub} that the communication complexity of the equality problem in this model is~$\Omega(\sqrt n)$, and therefore~$c$ must be~$\Omega(\sqrt{n})$. The third part follows from considering the~$k$-party equality function. The upper bound is shown in Claim~\ref{c_k-eq}. The lower bound follows from Newman and Szegedy~\cite{NS96_Pub}, because any~$k$-party SMP protocol without shared randomness among~$\A_1,\dots,\A_k$ for the~$k$-party equality function can be used to construct a two-party SMP protocol without shared randomness between two players~$\A'_1$ and~$\A'_2$ for the two-party equality function without affecting its cost: $\A'_1$ simulates~$\A_1$, and~$\A'_2$ simulates~$\A_2,\dots,\A_k$. \end{proof} \begin{remark} Besides its own elegance, the hierarchy of shared randomness is a useful technical tool. For our lower bound proof for the quantum complexity of~$\GPk$ (which is the main technical result of Section~\ref{s_vs}), we need different modes of shared randomness. Informally speaking, we use a hybrid argument that puts certain restrictions on the input values, and those restrictions inevitably create shared randomness of certain type that becomes available to the players. We show that the sort of randomness that is introduced corresponds to one of the restricted modes of shared randomness, whose availability does not make the communication problem easy for quantum communication. \end{remark} \section{Shared randomness vs.\ quantum communication} \label{s_vs} In this section, we will analyze the complexity of~$\GPk$ to compare the resource of shared randomness to that of quantum communication in multi-party protocols. Fix~$k\ge3$. Recall that in the classical SMP model with shared randomness~$\GPk$ has a trivial solution, where each player sends one bit to the referee. As a warm-up, consider the case of quantum protocols \emph{without shared randomness} \footnote We shall see soon why in the actual proof we have to consider different modes of shared randomness, even if our only purpose was to get a lower bound on the complexity of~$\GPk$ in the quantum model without shared randomness.} Let~$\Pk$ be a quantum protocol that communicates~$c$ qubits and solves~$\GPk$, and let~$\Uk$ be the uniform distribution over~$k$-tuples~$(x_1,\dots,x_k)\in\zo^{n\times k}$. Now consider the behavior of~$\Pk$ when the input is distributed according to~$\Uk$ (note that such input is almost never valid for~$\GPk$). For~$i\in[k]$, let~$\sigma_i$ be a density matrix representing the (mixed) state that the referee receives from~$\A_i$ when the input distribution is~$\Uk$. Since the senders share no randomness and~$\Uk$ is a product distribution, the state of the referee before his measurement is performed can be written as $\sigma_1\otimes\dots\otimes\sigma_k$. For~$i\in[k]$, let~$X_i$ be an~$n$-bit random string taking the value of input~$x_i$. By Claim~\ref{c_rac}, there exists~$j_0 \in [n]$ such that $\sum_{i=1}^k\norm{\sigma_0^i-\sigma_1^i}_1^2 \in O(c/n)$, where~$\sigma_a^i$ is the message from~$\A_i$, conditional upon~$X_i(j_0)=a$. Since the random variables~$X_1(j_0),\dots,X_k(j_0)$ are mutually independent and each~$X_i(j_0)$ can be correlated only with~$\sigma_i$, Claim~\ref{c_xor} implies that the referee can predict~$X_1(j_0)\oplus\dots\oplus X_k(j_0)$ with probability at most $1/2+(1/2^{k+1})\prod_{i=1}^k\norm{\sigma_0^i-\sigma_1^i}_1$, which is at most~$1/2+O(c/(kn))^{k/2}$ by the inequality of arithmetic and geometric means. Note that this guarantees that the ``advantage over random guess'' that the referee can have in predicting~$X_1(j_0)\oplus\dots\oplus X_k(j_0)$ using the messages received from the players is~$o(1/n)$, as long as~$k\ge3$ and~$c\in o(kn^{1-2/k})$. Moreover, similar reasoning can be applied to conclude that for most of the values of~$j_0\in[n]$, the possible advantage in predicting~$X_1(j_0)\oplus\dots\oplus X_k(j_0)$ must be very small. With this observation in hand, we would like to apply a ``hybrid-like'' reasoning, arguing that in order to distinguish between those inputs where most of bitwise XORs equal~$0$ and those where most equal~$1$, a protocol should be able, informally, to ``accumulate advantage'' from different input positions. We would like to claim that this is impossible as long as the advantage is negligible for almost every~$j_0\in[n]$. Here comes the main subtlety of our proof. Note that using hybrid-like argument puts a condition on a part of the input: specifically, in order for the ``hybrid scenario'' to get through, it has to be argued that it is hard for the protocol to predict most of the values of~$X_1(j)\oplus\dots\oplus X_k(j)$, even if the players ``know'' the values of~$X_1(j')\oplus\dots\oplus X_k(j')$ for those positions~$j'$ that were considered in the earlier stages of the induction. But such conditioning creates certain type of shared randomness between the players, and we can no longer assume mutual independence of the messages received by the referee, as we have done in the reasoning above. Recall that we are dealing with a communication problem that is easy in the presence of shared randomness even classically. How can we hope for quantum hardness, as required for the hybrid argument to be applicable? It turns out that \emph{the mode of shared randomness that results from using the hybrid method is not powerful enough to make the problem easy, even for quantum communication}. Our proof of Lemma~\ref{l_hyb-step} below follows rather closely the outline given above, but it also contains some new ingredients required to make the argument robust against weaker modes of shared randomness. \subsection{Exponential Separation for multi-party protocols} \label{ss_proof} We are ready to prove our main technical statement. \begin{reptheorem}{t_main} \theomain \end{reptheorem} First, we set up notation to describe protocol~$\Pk$ which uses~$t$-shared randomness. Let~$\Vkt=\{S \subseteq [k]\colon |S|=t\}$. For any~$S \in \Vkt$, let~$R_S$ be the random string (of arbitrary length) shared by the players~$\A_i$ for~$i \in S$. Then~$\A_i$ holds the~$R_S$'s for all~$S \in \Vkt$ containing~$i$. For convenience, let~$\tilde{R}_i=(R_S)_{i \in S \in \Vkt}$ be the~$\binom{k-1}{t-1}$-tuple of these random strings. For each instantiation~$R_S=r_S$, let~$\tilde{r}_i=(r_S)_{i \in S \in \Vkt}$ be the corresponding instantiation of~$\tilde{R}_i$. In addition, let~$\vR=(R_S)_{S \in \Vkt}$ be the~$\binom{k}{t}$-tuple of all shared random strings, and let~$\vr=(r_S)_{S \in \Vkt}$ be any instantiation of~$\vR$. For each~$i \in [k]$, player~$\A_i$ sends a quantum state~$\rho^i_{x_i,\tilde{r}_i}$ conditional upon receiving input~$x_i$ and random strings~$\tlR_i=\tlr_i$. Let~$c_i$ be the length of the quantum message sent by~$\A_i$; i.e., the length of state~$\rho^i_{x_i,\tlr_i}$ is~$c_i$ qubits. By assumption, $c \deq \sum_{i=1}^k c_i = o(kn^{1-t/k})$. As before, let~$\Uk$ be the uniform distribution over~$k$-tuples~$(x_1,\dots,x_k)\in\zo^{n\times k}$. To prove the theorem, we will use the following lemma. \begin{lemma} \label{l_hyb-step} Let~$2\le t\le k-1$, and let~$\Pk$ be a quantum SMP protocol of cost~$c=o(kn^{1-t/k})$ that uses~$t$-shared randomness as defined above. Suppose the player~$\A_i$ receives the random input~$X_i$ for~$(X_1,\dots,X_k) \in \U(\zonk)$. Then there exists~$J \subseteq [n]$ of size at least~$2n/3$ such that for every~$j \in J$ a referee who is allowed to apply an arbitrary quantum measurement to the messages received according to~$\Pk$ can predict the value of~$X_1(j)\oplus\dots\oplus X_k(j)$ with probability at most~$1/2+o(1/n)$. \end{lemma} \begin{proof} Let~$\sigma^i_{a,\vr}(j)$ be the expectation of~$\rho^i_{X_i,\tlR_i}$ conditional upon~$X_i(j)=a$ and~$\vR=\vr$, for any~$i \in [k]$, $j \in [n]$, $a \in \zo$ and possible~$\vr$. Define~$\alpha_{i,\vr} \in \mathbb{R}^n$ as \begin{equation} \alpha_{i,\vr}(j)=\frac12 \norm{\sigma^i_{0,\vr}(j)-\sigma^i_{1,\vr}(j)}_1. \label{eq:defaivj} \end{equation} Then by Claim~\ref{c_rac}, \[ \sum_{j=1}^n (\alpha_{i,\vr}(j))^2 \le O(c_i). \] Taking the sum of both sides over~$i \in [k]$ and using~$\sum_{i=1}^k c_i=c$ yields \[ \sum_{i=1}^k\sum_{j=1}^n (\alpha_{i,\vr}(j))^2 \le O(c). \] This holds for any possible~$\vr$. So \[ \E*{\sum_{i=1}^k\sum_{j=1}^n (\alpha_{i,\vR}(j))^2}\le O(c). \] (Here the expectation is taken with respect to the~$R_S$'s). Thus, there exists some~$J \subseteq [n]$ of size at least~$2n/3$ such that, for any~$j_0 \in J$, \begin{equation} \E*{\sum_{i=1}^k (\alpha_{i,\vR}(j_0))^2} \le O\left(\frac{c}{n}\right). \label{eq:sum_aivij} \end{equation} Let~$\sigma_{a,\vr}(j_0)$ be the expectation of~$\rho^1_{X_1,\tlR_1} \otimes \dots \otimes \rho^k_{X_k,\tlR_k}$ conditional upon~$X_1(j_0)\oplus \dots \oplus X_k(j_0)=a$ and~$\vR=\vr$, for any~$a \in \zo$ and possible~$\vr$. Since the~$X_i(j_0)$'s are i.i.d.\ with~$X_i(j_0)\sim \U(\zo)$, and they are also independent from the~$R_S$'s, we have \[ \sigma_{a,\vr}(j_0)=\E_{a_1 \oplus \dots \oplus a_k=a} {\sigma^1_{a_1,\tlr_1}(j_0) \otimes \dots \otimes \sigma^k_{a_k,\tlr_k}(j_0)}. \] (Here the expectation is taken with respect to the~$a_i$'s). So by Claim~\ref{c_xor} and~(\ref{eq:defaivj}) we get \[ \norm{\sigma_{0,\vr}(j_0)-\sigma_{1,\vr}(j_0)}_1 = 2 \prod_{i=1}^k \alpha_{i,\vr}(j_0). \] Now let~$\sigma_{a}(j_0)$ be the expectation of~$\rho^1_{X_1,\tlR_1} \otimes \dots \otimes \rho^k_{X_k,\tlR_k}$ conditional upon $X_1(j_0)\oplus \dots \oplus X_k(j_0)=a$, for~$a \in \zo$. Then~$\sigma_{a}(j_0)=\E{\sigma_{a,\vR}(j_0)}$. Thus, \begin{align*} \norm{\sigma_{0}(j_0)-\sigma_{1}(j_0)}_1 &=\norm{\E{\sigma_{0,\vR}(j_0)}-\E{\sigma_{1,\vR}(j_0)}}_1 \\ &\le \E{\norm{\sigma_{0,\vR}(j_0)-\sigma_{1,\vR}(j_0)}_1} \\ &= 2\E*{\prod_{i=1}^k \alpha_{i,\vR}(j_0)}. \end{align*} Note that~$Z_i\deq\alpha_{i,\tlR_i}(j_0)$ is a non-negative function of the $R_S$'s for~$i \in S \in \Vkt$. (Recall that~$\tlR_i=(R_S)_{i \in S \in \Vkt}$.) Since the~$R_S$'s are independent random variables, and each~$R_S$ is read~$t$ times (by the~$Z_i$'s for $i \in S$), we know that~$Z_1,\dots,Z_k$ form a read-$t$ family. Thus, by invoking Lemma~\ref{l_Hold}, we get \[ \E* {\prod_{i=1}^k \alpha_{i,\vR}(j_0)} \le \left(\prod_{i=1}^k \E{(\alpha_{i,\vR}(j_0))^t}\right)^{1/t}. \] Since~$0 \le \alpha_{i,\vR}(j_0) \le 1$ and~$t\ge2$, we have \[ \left(\prod_{i=1}^k \E{(\alpha_{i,\vR}(j_0))^t}\right)^{1/t} \le \left(\prod_{i=1}^k \E{(\alpha_{i,\vR}(j_0))^2}\right)^{1/t}. \] Then by the inequality of arithmetic and geometric means and~(\ref{eq:sum_aivij}), \begin{align*} {\left(\prod_{i=1}^k \E{(\alpha_{i,\vR}(j_0))^2}\right)\!\!}^{1/t} &\le {\left(\frac{1}{k}\sum_{i=1}^k \E{(\alpha_{i,\vR}(j_0))^2}\right)\!}^{k/t} \\ &\le \left( O\left( \frac{c}{kn}\right)\right)^{k/t} = o\left(\frac{1}{n}\right), \end{align*} provided~$c=o(kn^{1-t/k})$. So, we have~$\norm{\sigma_{0}(j_0)-\sigma_{1}(j_0)}_1=o(1/n)$. Namely, the referee can predict the value of~$X_1(j_0) \oplus \dots \oplus X_k(j_0)$ with probability at most~$1/2+o(1/n)$. This holds for any $j_0 \in J$. \end{proof} \begin{proof}[Proof of Theorem~\ref{t_main}] Let~$\Pk$ be a quantum SMP protocol of cost~$c=o(kn^{1-t/k})$ that uses~$t$-shared randomness. We will show that there exist~$L=\floor{3n/4}$ coordinate \footnote In fact, our statement holds for any~$L \le (1-\varepsilon)n$, where~$\varepsilon$ can be any small constant.} $j_1,\dots,j_L \in [n]$ satisfying the following conditions. For~$l \in [L]$ and~$a \in \zo$, let~$W_{l,a}$ be the set of~$k$-tuples~$(x_1,\dots,x_k) \in \zonk$ satisfying~$x_1(j) \oplus \dots \oplus x_k(j)=a$ for~$j=j_1,j_2,\dots,j_l$, and let~$\tau_{l,a}=\E{\rho^1_{X_1,\tlR_1}\otimes\dots \otimes \rho^k_{X_k,\tlR_k}}$ for~$(X_1,\dots,X_k) \sim \U(W_{l,a})$ (here the expectation is taken with respect to the~$X_i$'s and~$R_S$'s). Then: (i) $\norm{\tau_{1,0}-\tau_{1,1}}_1=o(1/n)$; (ii) for any~$l \in [L-1]$, $\norm{\tau_{l,0}-\tau_{l+1,0}}_1=o(1/n)$ and~$\norm{\tau_{l,1}-\tau_{l+1,1}}_1=o(1/n)$. If this is true, then by the triangle inequality, \begin{align} &\norm{\tau_{L,0}-\tau_{L,1}}_1 \nonumber \\ & \le \sum_{l=1}^{L-1}\norm{\tau_{l,0}-\tau_{l+1,0}}_1 +\sum_{l=1}^{L-1}\norm{\tau_{l,1}-\tau_{l+1,1}}_1 \nonumber \\ &\quad +\norm{\tau_{1,0}-\tau_{1,1}}_1 \nonumber \\ &=o(L/n) \nonumber \\ &=o(1). \label{eq:triseq} \end{align} On the other hand, for any~$(x_1,\dots,x_k) \in W_{L,0}$, it holds that~$|x_1 \oplus x_2 \oplus \dots \oplus x_k|\le n-L < n/3$ and hence~$\GPk(x_1,\dots,x_k)=0$. Similarly, for any~$(x_1,\dots,x_k) \in W_{L,1}$, it holds that~$|x_1 \oplus x_2 \oplus \dots \oplus x_k|\ge L > 2n/3$ and hence~$\GPk(x_1,\dots,x_k)=1$. Therefore, if the referee can correctly predict the value of~$\GPk(x_1,\dots,x_k)$ on any~$(x_1,\dots,x_k) \in W_{L,0} \sqcup W_{L,1}$, then he should be able to distinguish between~$\tau_{L,0}$ and~$\tau_{L,1}$ with probability at least $2/3$, which implies that~$\norm{\tau_{L,0}-\tau_{L,1}}_1=\Omega(1)$. But this is contradictory to~(\ref{eq:triseq}). So the referee must fail to solve~$\GPk$ on some valid input from~$W_{L,0}$ or~$W_{L,1}$. To find the desired~$j_1,\dots,j_L$, we use one initial step and~$L-1$ inductive steps as follows. \textbf{Initial Step:} Consider the behavior of~$\Pk$ on the random input~$(X_1,\dots,X_k) \in \U(\zonk)$. By a straightforward application of Lemma~\ref{l_hyb-step}, there exists~$J \subseteq [n]$ of size at least~$2n/3$ such that, for any~$j \in J$, the referee can predict the value of~$X_1(j) \oplus \dots \oplus X_k(j)$ with probability at most~$1/2+o(1/n)$. In other words, for any~$j \in J$, we have $\norm{\sigma_0(j)-\sigma_1(j)}_1=o(1/n)$ where~$\sigma_{a}(j)$ is the expectation of~$\rho^1_{X_1,\tlR_1} \otimes \dots \otimes \rho^k_{X_k,\tlR_k}$ conditional upon~$X_1(j)\oplus \dots \oplus X_k(j)=a$, for~$a \in \zo$. Set~$j_1$ to be any~$j \in J$. Then~$\norm{\tau_{1,0}-\tau_{1,1}}_1=\norm{\sigma_0(j)-\sigma_1(j)}_1=o(1/n)$ as desired. \textbf{Inductive Step:} Suppose now we have fixed~$j_1,\dots,j_l \in [n]$ for some~$l \le 3n/4$. Let~$T=\{j_1,\dots,j_l\}$ and~$T^c=[n] \setminus T$. Let us consider the behavior of~$\Pk$ on the random input~$(X_1,\dots,X_k) \sim \U(W_{l,0})$ (recall that~$W_{l,0}$ is the set of~$(x_1,\dots,x_k) \in \zonk$ satisfying~$(x_1 \oplus \dots \oplus x_k)_T=\bar{0}$). Note that the~$X_i$'s are not completely independent but only~$(k-1)$-wise independent. So there exists some correlation among the inputs to different players, which might be exploited to gain some advantage. However, we will show that, even in this case, the referee can still predict the value of~$X_1(j) \oplus \dots \oplus X_k(j)$ with probability at most~$1/2+o(1/n)$ for at least~$2/3$ fraction of~$j \in T^c$. Let~$Y_i=(X_i)_T$. Then~$(Y_1,\dots,Y_k)$ is uniformly distributed among all~$(y_1,\dots,y_k)\in \zolk$ satisfying~$y_1 \oplus \dots \oplus y_k=\bar{0}$. Namely, $(Y_1,\dots,Y_k)$ can be viewed as some XOR randomness (which is a special kind of~$2$-shared randomness) shared by the players. Also, note that the~$(X_i)_{T^c}$'s are i.i.d.\ with~$(X_i)_{T^c} \sim \U(\zo^{n-l})$, and they are also independent from the~$Y_i$'s. Finally, the~$Y_i$'s and~$(X_i)_{T^c}$'s are all independent from the~$R_S$'s. Now consider the following protocol~$\Pk'$ which attempts to solve~$\GPk$ for~$(n-l)$-bit strings. The players share XOR randomness~$(Y_1',\dots,Y_k')$ which has the same distribution as~$(Y_1,\dots,Y_k)$. In addition, they also share~$t$-shared randomness~$(R_S')_{S \in \Vkt}$ which has the same distribution as~$(R_S)_{S \in \Vkt}$. Furthermore, the~$Y_i'$'s and~$R_S'$'s are independent. Now suppose player~$\A_i$ receives input~$x_i' \in \zo^{n-l}$ and random strings~$Y_i'=y_i$ and~$\tlR_i'=\tlr_i$. Then~$\A_i$ first finds the unique~$x_i \in \zon$ such that~$(x_i)_T=y_i$ and~$(x_i)_{T^c}=x_i'$, and then sends the quantum message~$\rho^i_{x_i,\tlr_i}$ of~$\Pk$ to the referee. Since the~$(Y_1',\dots,Y_k')$ is a special kind of~$2$-shared randomness, $\Pk'$ uses only~$t$-shared randomness. So by Lemma~\ref{l_hyb-step} (replacing the original~$n$ by~$n-l$ and noting that~$n-l=\Theta(n)$, since~$l \le 3n/4$), we know that, on the random input~$(X_1',\dots,X_k') \sim \U(\zo^{(n-l)\times k})$, there exists~$J' \subseteq [n-l]$ of size at least~$2(n-l)/3$ such that for any~$j \in J'$ the referee can predict the value of~$X_1'(j)\oplus \dots \oplus X_k'(j)$ with probability at most~$1/2+o(1/n)$ using the messages received according to~$\Pk'$. Meanwhile, by the construction of~$X_i'$'s, $Y_i'$'s, $R_S'$'s and~$\Pk'$, it is obvious that the joint message sent according to~$\Pk'$ has the same distribution as $\rho^1_{X_1,\tlR_1}\otimes\dots \otimes\rho^k_{X_k,\tlR_k}$. In addition, the bits of~$X_i'$ are in one-to-one correspondence with the bits of~$(X_i)_{T^c}$. Thus, getting back to the original protocol~$\Pk$, we know that, on the random input~$(X_1,\dots,X_k) \sim \U(W_{l,0})$, there exists~$J_0 \subseteq T^c$ of size at least~$2(n-l)/3$ (corresponding to~$J'$) such that for any~$j \in J_0$ the referee can predict the value of~$X_1(j)\oplus \dots \oplus X_k(j)$ with probability at most~$1/2+o(1/n)$ using the messages received according to~$\Pk$. So for any~$j \in J_0$, we have~$\norm{\sigma_0(j)-\sigma_1(j)}_1=o(1/n)$, where $\sigma_a(j)$ is the expectation of $\rho^1_{X_1,\tlR_1}\otimes\dots \otimes\rho^k_{X_k,\tlR_k}$ conditional upon~$X_1(j)\oplus \dots \oplus X_k(j)=a$ for~$a \in \zo$. Now if we set~$j_{l+1}=j$, then depending on the value of~$x_1(j)\oplus \dots \oplus x_k(j)$, $W_{l,0}$ is split into to two equal-sized subsets: $W_{l+1,0}$ and~$W_{l,0}\setminus W_{l+1,0}$. It follows that~$\tau_{l,0}=(\sigma_0(j)+\sigma_1(j))/2$ and~$\tau_{l+1,0}=\sigma_0(j)$, and hence $\norm{\tau_{l,0}-\tau_{l+1,0}}_1 =\norm{\sigma_{0}(j)-\sigma_{1}(j)}_1/2=o(1/n)$. By a similar argument, we can also prove that, on the random input~$(X_1,\dots,X_k) \sim \U(W_{l,1})$, there also exists~$J_1 \subseteq T^c$ of size at least~$2(n-l)/3$ such that for any~$j \in J_1$ the referee can predict~$X_1(j) \oplus \dots \oplus X_k(j)$ with probability at most~$1/2+o(1/n)$. Then, if we set~$j_{l+1}$ to be any~$j \in J_1$, then we get~$\norm{\tau_{l,1}-\tau_{l+1,1}}_1=o(1/n)$. Now since~$|J_0|,|J_1| \ge 2(n-l)/3$, $J_0 \cap J_1$ must be non-empty. We set~$j_{l+1}$ to be any~$j \in T_0 \cap T_1$. Then we achieve both $\norm{\tau_{l,0}-\tau_{l+1,0}}_1=o(1/n)$ and $\norm{\tau_{l,1}-\tau_{l+1,1}}_1=o(1/n)$. Iterate this inductive step~$L-1$ times, and in the end we obtain the desired~$j_1,\dots,j_L$. This completes a proof of the theorem. \end{proof} \section*{Acknowledgments} The authors thank anonymous reviewers for helpful comments on previous versions of this paper. The authors acknowledge support by ARO/NSA under grant W911NF-09-1-0569.
\section{Introduction} The interaction of a point charge and a magnetic moment represents one of the persistent problems of classical electromagnetism.\cite{Griffiths}$^{, $\cite{Jackson} \ The interaction is intriguing partly because the combination of the electric field of the charge and the magnetic field of the magnetic moment introduces electromagnetic field linear momentum. \ One mystery of the charge-magnetic-moment interaction is just how Nature incorporates this electromagnetic field momentum into the relativistic conservation law for linear momentum. \ The interaction of a point charge and a magnetic moment forms the basis for controversies involving \textquotedblleft hidden momentum,\textquotedblright\cite{Shockley}$^{,}$\cite{Vaidman}$^{, $\cite{RLetter} the Aharonov-Bohm effect,\cite{AB} and the Aharonov-Casher effect.\cite{AC} \ However, a certain version of the interaction involving hidden momentum has made its way into the textbook literature\cite{Griffiths2 $^{,}$\cite{Jackson2} despite some objections.\cite{objection} \ Indeed some research journals have rejected a competing description of the interaction, even refusing to send out for review manuscripts which explore the competing description.\cite{refuse} \ \ Most recently, the interaction has been used as the basis for the astonishing claim that the Lorentz force law is incompatible with relativity.\cite{Mansuripur} \ Given the controversies which still exist, it seems wise to review the varying suggestions for the interaction of a point charge and a magnetic moment. \ Here we will carry out calculations involving two different models for a magnetic moment which interact with a point charge in strikingly different ways. The contrasting behaviors suggest that there is yet more to be understood regarding the interaction as it exists in Nature. \ The contrast also suggests the possibility that the current \textquotedblleft hidden momentum\textquotedblright\ interpretation in the classical electromagnetism textbooks may be suspect and that the Aharonov-Bohm effect and Aharonov-Casher effect may be wrongly presented in the quantum mechanics texts. \section{Two Models for a Magnetic Moment} In the present article, two different models for a magnetic moment are considered. \ The unperturbed magnetic-moment models both have a point charge $e$ of mass $m$ moving in a circular orbit $\mathbf{r(}t)$ of radius $r_{0}$ and angular frequency $\omega_{0}$ about a charge $-e$ of large mass $M$ locate at the center of the orbit. \ Thus for the unperturbed magnetic moment orbit, we have the displacement, velocity, and acceleration of the charge $e$ located at angle $\phi=\omega_{0}t+\phi_{0}$ given b \begin{equation} \mathbf{r(}t)=r_{0}[\widehat{i}\cos(\omega_{0}t+\phi_{0})+\widehat{j \sin(\omega_{0}t+\phi_{0})] \label{ee1 \end{equation \begin{equation} \mathbf{v}(t)=\omega_{0}r_{0}[-\widehat{i}\sin(\omega_{0}t+\phi_{0 )+\widehat{j}\cos(\omega_{0}t+\phi_{0})] \label{ee2 \end{equation \begin{equation} \mathbf{a}(t)=-\omega_{0}^{2}r_{0}[\widehat{i}\cos(\omega_{0}t+\phi _{0})+\widehat{j}\sin(\omega_{0}t+\phi_{0})] \label{ee3 \end{equation} The system has a magnetic moment \begin{equation} \overrightarrow{\mu}=\widehat{k}\frac{e\omega_{0}r_{0}^{2}}{2c} \label{ma1 \end{equation} obtained from the ensemble-averaged (averaging over the initial phase $\phi_{0})$ or time-averaged current density, and the steady-state current formula $\overrightarrow{\mu} {\displaystyle\int} d^{3}r\,\mathbf{r}\times\mathbf{J}/(2c).$\cite{Jackson3} \ \ The crucial difference between the models involves the binding which leads to the circular orbit. \ The magnetic-moment model favored by the proponents of hidden momentum involves a \textquotedblleft fixed-path\textquotedblright\ constraint such that during any interaction the point charge $e$ may accelerate along the circular orbital path but cannot depart from the circular orbital path of radius $r_{0}$. \ The magnetic-moment model favored by the present author involves a Coulomb-potential interaction between the charge $e$ and the central charge $-e,$ corresponding to the behavior of a hydrogen-like atom. \ Both these models have been presented previously in the literature, but never before in direct comparison.\cite{comparison} \section{Electromagnetic Momentum of Interaction} We assume that the external charge $q$ (with which the magnetic moment is interacting) is located on the $x$-axis at coordinate $x_{q},$ $\mathbf{r _{q}=\widehat{i}x_{q},$ where $x_{q}>>r_{0},$ so that we may think of the electric field $\mathbf{E}_{q}(\mathbf{r})=q(\mathbf{r}-\widehat{i x_{q})/|\mathbf{r}-\widehat{i}x_{q}|^{3}$ due to the charge $q$ as approximately constant over the magnetic moment at the value $\mathbf{E _{q}=-\widehat{i}q/x_{q}^{2}$ with small corrections of order $r_{0}/x_{q}.$ \ The location of the magnetic moment in the electric field of the charge $q$ leads to linear momentum {\displaystyle\int} d^{3}r\mathbf{E}\times\mathbf{B}/(4\pi c)$ in the electromagnetic fields in vacuum arising from the electric field $\mathbf{E}_{q}$ of the charge $q$ and the magnetic field $\mathbf{B}_{e}$ of the magnetic moment. \ The electromagnetic field momentum appears in order $v^{2}/c^{2},$ and therefore we can work with the Darwin Lagrangian\cite{Darwin} and calculate perturbations using nonrelativistic physics. \ Since the electromagnetic field linear momentum $\mathbf{p}_{em}$ is already first order in the perturbing charge $q$, the momentum\cite{PageandAdams} can be calculated through first order in $q$ by using the unperturbed motion of the charge $e$ a \begin{align} \mathbf{p}_{em} & =\frac{eq}{2c^{2}}\left( \frac{\mathbf{v}(t) {|\mathbf{r(t)-}\widehat{i}x_{q}|}+\frac{\mathbf{v}(t)\cdot(\mathbf{r(t)- \widehat{i}x_{q})(\mathbf{r(t)-}\widehat{i}x_{q})}{|\mathbf{r(t)- \widehat{i}x_{q}|^{3}}\right) \nonumber\\ & =\frac{eq}{2c^{2}}\left[ \frac{\mathbf{v}(t)}{x_{q}}\left( 1+\frac {\widehat{i}\cdot\mathbf{r}(t)}{x_{q}}\right) +\frac{\mathbf{v (t)\cdot\lbrack\mathbf{r(t)-}\widehat{i}x_{q}][\mathbf{r(t)-}\widehat{i x_{q}]}{x_{q}{}^{3}}\left( 1+3\frac{\widehat{i}\cdot\mathbf{r}(t)}{x_{q }\right) \right] \label{pp1 \end{align} where we have expanded the denominators assuming that $r_{0}<<x_{q}$ to obtain the expressions $|\mathbf{r(t)-}\widehat{i}x_{q}|^{-1}=x_{q}^{-1 (1+\widehat{i}\cdot\mathbf{r}(t)/x_{q}+...),$ $|\mathbf{r(t)-}\widehat{i x_{q}|^{-3}=x_{q}^{-3}(1+3\widehat{i}\cdot\mathbf{r}(t)/x_{q}+...).$ \ Then the average linear momentum $\left\langle \mathbf{p}_{em}\right\rangle $ in the electromagnetic field is found from Eqs. (\ref{ee1}) and (\ref{ee2}) by either ensemble averaging or averaging over time $t$ with $\left\langle \cos^{2}\theta\right\rangle =\left\langle \sin^{2}\theta\right\rangle =1/2,$ $\left\langle \cos\theta\sin\theta\right\rangle =0, \begin{equation} \left\langle \mathbf{p}_{em}\right\rangle =\frac{eq}{2c^{2}}\left( \widehat{j}\frac{\mathbf{\omega}_{0}r_{0}^{2}}{x_{q}^{2}}\right) =-\frac {1}{c}\overrightarrow{\mu}\times\mathbf{E} \label{pp2 \end{equation} This is the electromagnetic field momentum which must be consistent with the conservation of linear momentum. \section{Nonrelativistic Perturbation Calculation -- Fixed-Path Model} In the fixed-path model\cite{comparison} for a magnetic moment, the current-carrying charge $e$ remains on the circular path of radius $r_{0}$ but may change its velocity along this path. \ In the nonrelativistic perturbation calculation, the angular acceleration $d^{2}\phi/dt^{2}$ of the charge $e$ in its circular orbit is produced by the component of the electric field $\mathbf{E}_{q}$ tangential to the orbi \begin{equation} \frac{d^{2}\phi}{dt^{2}}=\frac{a_{\phi}}{r_{0}}=\frac{eE_{q}\sin(\phi) {r_{0}m}=\frac{eq\sin(\phi)}{mr_{0}x_{q}^{2}} \label{ee4 \end{equation} Since the angular acceleration is regarded as small, we may introduce the unperturbed expression $\phi_{unperturbed}(t)=\omega_{0}t+\phi_{0}$ on the right-hand side of Eq. (\ref{ee4}) to obtai \begin{equation} \frac{d^{2}\phi}{dt^{2}}=\frac{eq\sin(\omega_{0}t+\phi_{0})}{mr_{0}x_{q}^{2}} \label{ee5 \end{equation} Assuming that the unperturbed motion holds at time $t=0,$ this equation can be integrated with respect to time to obtai \begin{equation} \phi(t)=\omega_{0}t+\phi_{0}-\frac{eq\sin(\omega_{0}t+\phi_{0})}{\omega _{0}^{2}mr_{0}x_{q}^{2}} \label{ee6a \end{equation} Thus when interacting with the charge $q,$ the orbiting particle $e$ of the magnetic moment has the position \begin{align} \mathbf{r}(t) & =r_{0}[\widehat{i}\cos(\phi)+\widehat{j}\sin(\phi )]\nonumber\\ & =r_{0}\left[ \widehat{i}\cos\left( \omega_{0}t+\phi_{0}-\frac {eq\sin(\omega_{0}t+\phi_{0})}{\omega_{0}^{2}mr_{0}x_{q}^{2}}\right) +\widehat{j}\sin\left( \omega_{0}t+\phi_{0}-\frac{eq\sin(\omega_{0}t+\phi _{0})}{\omega_{0}^{2}mr_{0}x_{q}^{2}}\right) \right] \nonumber\\ & =r_{0}\widehat{i}\left[ \cos\left( \omega_{0}t+\phi_{0}\right) +\frac{eq\sin(\omega_{0}t+\phi_{0})}{\omega_{0}^{2}mr_{0}x_{q}^{2}}\sin (\omega_{0}t+\phi_{0})\right] \nonumber\\ & +r_{0}\widehat{j}\left[ \sin(\omega_{0}t+\phi_{0})-\frac{eq\sin(\omega _{0}t+\phi_{0})}{\omega_{0}^{2}mr_{0}x_{q}^{2}}\cos\left( \omega_{0 t+\phi_{0}\right) \right] \label{ee7 \end{align} where we have used the approximations for small $\delta,$ $\sin(\theta +\delta)\approx\sin\theta+\delta\cos\theta$ and $\cos(\theta+\delta )\approx\cos\theta-\delta\sin\theta.$ \ The first thing which we notice is that the interaction of the magnetic moment (fixed-path model) has led to a nonrelativistic (zero-order in $v/c)$ average electric dipole moment $\left\langle \overrightarrow{\mathfrak{p }\right\rangle .$\ \ Thus the average electric dipole moment follows from Eq. (\ref{ee7}) a \begin{equation} \left\langle \overrightarrow{\mathfrak{p}}\right\rangle =\left\langle e\mathbf{r}(t)\right\rangle =\widehat{i}\frac{e^{2}q}{2m\omega_{0}^{2 x_{q}^{2}}=-\frac{e^{2}}{2m\omega_{0}^{2}}\mathbf{E}_{q} \label{ee8 \end{equation} This looks rather like the polarization found for a charged harmonic oscillator in the electrostatic field $\mathbf{E}_{q}$ of a charge $q$ except that the polarization is in the \textit{opposite} direction from the polarizing electric field. \ Rather than the usual attraction between a charge and a polarizable material, we find here a repulsion. There is no average nonrelativistic linear momentum in the circular orbit of the charge $e$ since $\left\langle \mathbf{p}_{mech\text{ nonrel }\right\rangle =m\left\langle \mathbf{v}\right\rangle =m\left\langle d\mathbf{r}/dt\right\rangle =0$ using $d\mathbf{r}/dt$ from Eq. (\ref{ee7}). \ However, there is a net linear momentum at the relativistic order $v^{2}/c^{2}.$ \ The relativistic expression for mechanical linear momentum of a particle is $\mathbf{p}_{mech}=m\gamma\mathbf{v}=m\gamma c^{2 \mathbf{v}/c^{2}$ where we recognize $m\gamma c^{2}$ as the mechanical energy (rest energy plus kinetic energy) of the particle. \ The energy conservation law which goes along with the fixed-path perturbation analysis gives energy balance for mechanical plus electrostatic potential energy as \begin{equation} m\gamma_{0}c^{2}+\frac{eqr_{0}\cos\phi_{0}}{x_{q}^{2}}=m\gamma c^{2 +\frac{eqr_{0}\cos\phi}{x_{q}^{2}} \label{ee9 \end{equation} Thus we hav \begin{equation} \mathbf{p}_{mech}=m\gamma\mathbf{v}=\frac{m\gamma c^{2}\mathbf{v}}{c^{2 }=\left( m\gamma_{0}c^{2}+\frac{eqr_{0}\cos\phi_{0}}{x_{q}^{2}}-\frac {eqr_{0}\cos\phi}{x_{q}^{2}}\right) \frac{1}{c^{2}}\frac{d\mathbf{r}}{dt} \label{ee10 \end{equation} Then time-averaging using $d\mathbf{r}/dt$ from Eq. (\ref{ee7}) (or indeed from the unperturbed $\mathbf{v(}t)$ of Eq. (\ref{ee2})), we fin \begin{equation} \left\langle \mathbf{p}_{mech}\right\rangle =-\widehat{j}\frac{eq\omega _{0}r_{0}^{2}}{2x_{q}^{2}c^{2}} \label{ee11 \end{equation} This relativistic mechanical linear momentum is exactly equal in magnitude and opposite in direction from the electromagnetic linear momentum found in Eq. (\ref{pp2}). \ Thus it is claimed that the mechanical hidden momentum balances the electromagnetic field momentum giving a self-consistent, self-contained system of zero linear momentum with no mystery regarding conservation of linear momentum. \ Of course, this description of the charge-magnetic-moment interaction says nothing about the source of the crucial nonrelativistic (zero-order in $v/c$) forces which constrain the motion of the current-carrying charge $e$ so that it moves only in a circular path. \ Also, there is no mention of the unusual electrostatic force between the magnetic moment and the charge $q$ associated with the nonrelativistic electric dipole moment in Eq. (\ref{ee8}). \ In the opinion of the present writer, this fixed-path charge-magnetic-moment interaction description has no more validity than the interaction description of two point charges $e$ and $q$ which are at rest at a separation $x_{m}$ and are noted with triumph to have zero linear momentum. The interaction description remains worthlessly incomplete unless one accounts for the external nonelectromagnetic forces which are required for equilibrium or else (if there are no non-electromagnetic forces) one notes the time evolution of the system under the electromagnetic forces. \section{Nonrelativistic Perturbation Calculation -- Purely Electromagnetic Model} Our second model\cite{comparison} for a magnetic moment involves the charge $e$ attracted to the massive opposite charge $-e$ by Coulomb attraction. \ In this case, all the forces are electromagnetic, and there are no additional nonelectromagnetic forces of constraint which keep the charge $e$ in a circular orbit. \ Indeed, as pointed out by Solem\cite{Solem} in his article, \textquotedblleft The Strange Polarization of the Classical Atom,\textquotedblright\ the interaction of the nonrelativistic hydrogen atom with an external electric field $\mathbf{E}_{q}$ produces an electric dipole moment for the magnetic moment which is \textit{perpendicular} to the electric field $\mathbf{E}_{q}$. \ Qualitatively, the situation is easy to understand. \ A charge $e$ in a circular Coulomb orbit will be slowed down when moving toward the external charge $q,$ and therefore the charge $e$ will tend to fall in closer to the central charge $-e;$ on the other hand, the charge $e$ will be speeded up when moving away from the external charge $q,$ and therefore the charge $e$ will tend to move further away from the central charge $-e.$ \ Since the external charge $q$ produces only a small orbital perturbation of the charge $e$, we expect that the orbit of the charge $e$ will be an elliptical Coulomb orbit. \ Furthermore, the semi-major axis of the elliptical orbit will remain essentially unchanged since the charge $e$ is in approximately periodic motion, moving repeatedly towards and away from the external charge $q,$ successively losing and gaining kinetic energy from the field $\mathbf{E}_{q}.$ \ The behavior involving a changing elliptical Coulomb orbit of fixed semi-major axis is totally different from the tangential-velocity-changing behavior portrayed in the fixed-path analysis discussed above. \ The changing ellipticity of the Coulomb orbit leads to a changing average electric dipole field back at the external charge $q$ which produces forces on $q$ which are completely different from those predicted by the fixed-path analysis. The perturbation analysis of the Coulomb orbit by an external electric field has been given in other publications.\cite{hydro} \ Here we will sketch the analysis. \ The displacement $\mathbf{r}$ of the charge $e$ with mass $m$ in a Coulomb orbit with a charge $-e$ at the origin is given by\cite{check} \ \begin{equation} \mathbf{r}=\frac{3}{2}\frac{\mathbf{K}}{(-2mH_{0})^{1/2}}+\frac{1}{4H_{0 }\frac{d}{dt}[m(\mathbf{r}\times\mathbf{v}\times\mathbf{r}+m\mathbf{v}r^{2}] \label{kk1 \end{equation} where $\mathbf{K}$ is the Laplace-Runge-Lenz vector\cite{Goldstein \begin{equation} \mathbf{K}=\frac{1}{(-2mH_{0})^{1/2}}\left( [\mathbf{r}\times(m\mathbf{v )]\times(m\mathbf{v})+me^{2}\frac{\mathbf{r}}{r}\right) \label{kk2 \end{equation} and $H_{0}$ is the energy of the particl \begin{equation} H_{0}=\frac{1}{2}mv^{2}-\frac{e^{2}}{r} \label{kk3 \end{equation} The Laplace-Runge-Lenz vector $\mathbf{K}$ is constant in time for a Coulomb orbit. \ Thus the time-average value of $\mathbf{r}$ averaged over the orbit follows from Eq. (\ref{kk1}) as \begin{equation} \left\langle \mathbf{r}\right\rangle =\frac{3}{2}\frac{\mathbf{K} {(-2mH_{0})^{1/2}} \label{kk4r \end{equation} The nonrelativistic equation of motion for the charge $e$ in the presence of the perturbing electric field $\mathbf{E}_{q}$ is \begin{equation} m\frac{d^{2}\mathbf{r}}{dt^{2}}=-\frac{e^{2}\mathbf{r}}{r^{3}}+e\mathbf{E}_{q} \label{kk5 \end{equation} Then the time-derivative of $\mathbf{K}$ follows from Eqs. (\ref{kk2}) and (\ref{kk5}) as \begin{equation} \frac{d\mathbf{K}}{dt}=me[-2\mathbf{r}(\mathbf{v}\cdot\mathbf{E _{q})+\mathbf{E}_{q}(\mathbf{r}\cdot\mathbf{v})+\mathbf{v}(\mathbf{r \cdot\mathbf{E}_{q})] \label{kk6 \end{equation} Since the right-hand side of Eq. (\ref{kk6}) is already first order in the perturbing field $\mathbf{E}_{q}$, we may evaluate the time derivative through first order by averaging the right-hand side over an unperturbed Coulomb orbit to obtai \begin{equation} \frac{d\mathbf{K}}{dt}=\frac{3}{2}me[(\mathbf{r}\times\mathbf{v )\times\mathbf{E}_{q}=\frac{3}{2}e\mathbf{L}\times\mathbf{E}_{q =3cm\overrightarrow{\mu}\times\mathbf{E}_{q} \label{kk7 \end{equation} where $\mathbf{L}=m\mathbf{r}\times\mathbf{v}$ is the orbital angular momentum of the charge $e.~\ \ $Thus in the Coulomb-orbit model for a magnetic moment, the current-carrying charge $e$ may start out in a circular orbit where the Laplace-Runge-Lenz vector $\mathbf{K}$ is zero, but the orbit changes in time becoming increasingly elliptical. \ The magnetic moment $\overrightarrow{\mu}$ changes in time along with the orbital angular momentum $\mathbf{L} \begin{align} \left\langle \frac{d\overrightarrow{\mu}}{dt}\right\rangle & =\frac{e {2mc}\left\langle \frac{d\mathbf{L}}{dt}\right\rangle =\frac{e}{2mc \left\langle \overrightarrow{\mathbf{\Gamma}}\right\rangle =\frac{e {2mc}\left\langle e\mathbf{r}\right\rangle \times\mathbf{E}_{q}\nonumber\\ & =\frac{e^{2}}{2mc}\frac{3}{2}\frac{\mathbf{K\times E}_{q}}{(-2mH_{0 )^{1/2}} \label{kk8 \end{align} The changing orbit of the charge $e$ in the magnetic moment will lead to electrical forces back on the external charge $q,$ both nonrelativistic electrostatic forces associated with the presence of the electric dipole moment and also relativistic forces associate with the electric field induced by the changing magnetic moment. \ It has been pointed out that the forces associated with this changing magnetic moment are qualitatively appropriate to account for the Aharonov-Bohm phase shift as a lag effect associated with classical electromagnetic forces.\cite{b2006} \subsection{Closing Summary} The problem of the interaction of a point charge and a magnetic moment is an old problem surrounded by controversy. \ In recent years, one version (involving hidden momentum) of this controversy has been introduced into the textbook literature of electromagnetism. \ In this article, we present two models for magnetic moments and note their contrasting behaviors in the presence of an external charged particle. \ On the one hand, the fixed-path model for the magnetic moment indeed exhibits hidden momentum, but it involves unmentioned nonelectromagnetic forces which are of nonrelativistic order and are vastly greater than the relativistic mechanical effects which are touted in the textbook literature. \ The fixed-path model also involves an unusual nonrelativistic electric dipole moment. \ On the other hand, the Coulomb-orbit model involves only electromagnetic interactions. \ This electromagnetic model shows a changing magnetic moment which introduces both electrostatic fields and induced electric fields. \ We believe that the interaction of a charged particle and a magnetic moment (appropriate for describing Nature) remains a poorly-understood aspect of electromagnetic theory and that it is premature to accept the hidden-momentum description of the interaction. \bigskip \bigskip
\section{Introduction} \setcounter{page}{1} Following the growing criticism of elliptical models, copulas have emerged both in insurance and risk management as a powerful alternative for modeling the complete dependence structure of a multivariate distribution. Since the seminal work by \cite{embrechts:2002}, the literature on copulas and their use in risk management applications has grown exponentially with several studies concentrating on statistical inference and model selection for copulas \citep[see, e.g.,][]{kim:2007,genestelal:2009} as well as applications \citep[see, e.g.,][]{Chan,grundke:2012,yea}.\footnote{A literature review with a special emphasis on finance-related papers using copulas is given by \cite{genest:2009}. An overview of the different branches of the copula literature is given by Embrechts (2009).} As simple parametric models are often not flexible enough to model high-dimensional distributions, recent works by \cite{joe:1996, joe:1997}, \cite{bedford:2001,bedford:2002} and \cite{whelan:2004} have proposed copula models which are highly flexible but at the same time still tractable even in higher dimensions. Most notably, vine copulas (also called pair-copula constructions, PCC in short) have emerged as the most promising tool for modeling dependence structures in high dimensions.\footnote{Competing modeling concepts like nested and hierarchical Archimedean copulas are analyzed by \cite{aas:2009b} as well as \cite{fischer:2009}. They conjecture that vine copulas should be preferred over nested or hierarchical Archimedean copulas.} Vine copulas consist of a cascade of conditional bivariate copulas (so called pair-copulas) which can each be chosen from a different parametric copula family. As a result, vine copulas are extremely flexible yet still tractable even in high dimensions as all computations necessary in statistical inference are performed on bivariate data sets \citep[see][for a first discussion of vine copulas in an applied setting]{aas:2009a}. Similar to the bivariate case,\footnote{See \citet{genestelal:2009}, \citet{kole:2007} and \citet{weiss:2011,weiss:2012rqfa} for discussions of the problem of selecting the best fitting parametric copula.}\ the correct selection of the parametric constituents of the vine, i.e., the pair-copulas, is crucial for the correct specification of a vine copula model. In case of the popular C- and D-vine specifications, the calibration and estimation of a $d$-dimensional vine requires the selection and estimation of $d(d-1)/2$ different pair-copulas from the set of candidate bivariate parametric copula families. Thus, a vine model's increased flexibility only comes at the expense of an increased model risk. As a remedy, recent studies have suggested to select the parametric pair-copulas based on graphical data inspection and goodness-of-fit tests \citep{aas:2009a} and to employ sequential heuristics based on Akaike's Information Criterion (AIC) \citep[see, e.g.,][]{brech:2012,dissmann:2011}. \citet{kurowicka:2010} and \citet{brech:2012} propose strategies for simplifying vines by replacing certain pair-copulas by the independence copula (yielding a \textit{truncated} vine copula) or the Gaussian copula (yielding a \textit{simplified vine}). Finally, \citet{hobaekhaff:2012} propose the use of empirical pair-copulas in vine models to circumvent the problem of selecting parametric pair-copulas. In this paper, we use the recently proposed nonparametric Bernstein copulas \citep{sancetta:2004,pfeifer:2009,diers:2012 \ as pair-copulas yielding smooth nonparametric vine copula models that do not require the specification of parametric families. Thus, we extend the ideas laid out by \citet{hobaekhaff:2012} by using an approximation to the empirical pair-copulas. In contrast to their work, however, we approximate the pair-copulas not only nonparametrically but also by the use of continuous functions.\footnote{Using smooth functions to approximate the true underlying dependence structure is in line with our intuition. The superiority of smooth nonparametric approximations of the copula over simple empirical copulas, however, is also found by \citet{shen:2008}. They argue that the improved approximation by the linear B-spline copulas is due to their Lipschitz continuity.}\ In addition, especially the Bernstein copula has recently attracted attention in insurance modeling \citep{diers:2012} and has already proven its merits in an applied setting. Therefore, the contributions of the proposed smooth nonparametric vine copulas are twofold: First, the use of Bernstein \ copulas completely obviates the need for the error-prone selection of pair-copulas from pre-specified sets of parametric copulas. The resulting smooth and nonparametric vine copulas do not only constitute extremely flexible tools for modeling high-dimensional dependence structures, they are also characterized by a smaller model risk than their parametric counterparts. Second, Bernstei \ copulas have been shown to improve on the estimation of the underlying dependence structure by competing nonparametric empirical copulas.\footnote{For example, Bernstein copulas provide a higher rate of consistency than other common nonparametric estimators and do not suffer from boundary bias \citep{kulpa:1999,sancetta:2004,diers:2012}. Similarly, other approximations as, e.g., linear B-spline copulas have also been shown to yield lower average squared approximation errors than competing discrete approximations \citep{shen:2008}.}\ The modeling of a vine model's pair-copulas by the use of smooth approximating functions is thus a natural extension of recently proposed (highly discontinuous) empirical pair-copulas. The usefulness of the proposed Bernstei \ vines is demonstrated by means of a simulation study as well as by forecasting and backtesting the Value-at-Risk (VaR) for multivariate portfolios of financial assets. The results presented in this study show that our proposed vine copula model with smooth nonparametric Bernstein pair-copulas outperforms the benchmark model with parametric pair-copulas in higher dimensions with respect to the accuracy and numerical stability of the approximation to the true underlying dependence structure. While our nonparametric vine copula model yields worse average squared errors than a benchmark vine copula calibrated by selecting parametric pair-copulas based on AIC values in lower dimensions (e.g., $d=3,5,7$) in our simulations, this result is reversed in higher dimensions. For random vectors of dimension $d=11$ and higher, the parametric benchmark broke down in more than $50\%$ of the simulations due to either the numerical instability of the parameter and AIC estimation or simply due to extremely inaccurate approximations caused by badly selected parametric pair-copulas. In higher dimensions (i.e., the main field of application of vine copulas), our nonparametric modeling approach is thus clearly superior to a parametric vine copula model. Our risk management application, however, shows that even in lower dimensions ($d=5$) our nonparametric model yields VaR forecasts that are not rejected by a range of formal statistical backtests. Consequently, the slightly worse approximation errors of our nonparametric model in lower dimensions do not seem to affect the modeling of a given dependence structure too severely thus underlining the usefulness of our proposed model. The remainder of this article is structured as follows. Section 2 introduces vine copulas as well as Bernstein copulas which we employ as pair-copulas in the vines. Section 3 presents the results of a simulation study on the approximation errors of both our nonparametric Bernstein vine copula model as well as a heuristically calibrated parametric benchmark model. In Section 4, we conduct an empirical analysis for a five-dimensional financial portfolio. Concluding remarks are given in Section 5. \section{Smooth nonparametric vine copulas}\label{sec:firstpart} The purpose of this section is to shortly introduce the fundamentals of vine and Bernstein copulas \subsection{Vine copulas} Vine copulas are obtained by hierarchical cascades of conditional bivariate copulas and are characterized by an increased flexibility for modeling inhomogeneous dependence structures in high dimensions. Here, we only present the basic definition as well as some basic properties of two popular classes of vines, i.e., C- and D-vines,\footnote{The classes of C- and D-vines are subsets of the so-called regular vines \citep[or R-vines in short, see][]{brech:2012}. We do not consider other types of R-vines in this paper but note that our proposed use of smooth Bernstein and B-spline copulas as pair-copulas can also be extended to other subsets of R-vines.} which will be used later on in our application to financial market data. Readers interested in a more rigorous examination of vine copulas and their properties are referred to the excellent studies by \citet{joe:1996,joe:1997} and \citet{bedford:2001,bedford:2002}. Starting point is the well-known observation that a joint probability density function of dimension $d$ can be decomposed into \begin{equation}\label{decomposed-density} f(x_1,\ldots,x_d)=f(x_1)\cdot f(x_2|x_1)\cdot f(x_3|x_1,x_2)\cdot\ldots\cdot f(x_d|x_1,\ldots,x_{d-1}). \end{equation} Each factor in this product can then be decomposed further using a conditional copula, e.g. \begin{equation}\label{decomposed-density2} f(x_2|x_1)=c_{12}(F_1(x_1),F_2(x_2))\cdot f_2(x_2) \end{equation} with $F_i(\cdot)$ being the cumulative distribution function (cdf) of $x_i$ ($i=1,\ldots,d$) and $c_{12}(\cdot)$ being the (in this case unconditional) copula density of $(x_1,x_2)$. Going further down the initial decomposition, the conditional density $f(x_3|x_1,x_2)$ could be factorised via \begin{equation}\label{decomposed-conddensity} f(x_3|x_1,x_2)=c_{23|1}(F_{2|1}(x_2|x_1),F_{3|1}(x_3|x_1))\cdot c_{13}(F_1(x_1),F_3(x_3))\cdot f_3(x_3) \end{equation} with $c_{23|1}(\cdot)$ being the conditional copula of $(x_2,x_3)$ given $x_1$. Finally, substituting the elements of the initial decomposition in \eqref{decomposed-density} with the conditional copulas, we get for dimension $d=3$ \begin{eqnarray}\label{3dvine} f(x_1,x_2,x_3)&=&c_{23|1}(F_{2|1}(x_2|x_1),F_{3|1}(x_3|x_1))\\ &\cdot& c_{12}(F_1(x_1),F_2(x_2))\notag\\ &\cdot& c_{13}(F_1(x_1),F_3(x_3))\notag\\ &\cdot& f_1(x_1)\cdot f_2(x_2)\cdot f_3(x_3)\notag \end{eqnarray} with $c_{12}$, $c_{13}$ and $c_{23|1}$ as \textit{pair-copulas}. Note that as there are several possible decompositions of the conditional distributions, the joint density of $x_1,\ldots,x_d$ can also be represented by different vine decompositions depending on the variables one chooses to condition on. \citet{bedford:2001,bedford:2002} propose representing these decompositions of a $d$-dimensional joint density as a nested set of trees where two nodes are joined by an edge in tree $j+1, j=1,\ldots,d-1,$ only if the corresponding edges in tree $j$ share a common node. Consequently, there are $d-1$ trees, where tree $j$ has $d+1-j$ nodes and $d-j$ edges with each edge corresponding to a pair-copula density, i.e., a density of a conditional bivariate parametric copula. In a Canonical or C-vine, each tree has a unique node (without loss of generality this is node $1$) that is connected to all other nodes yielding the representation \begin{equation}\label{decomposed-density c-vine} f(x_1,\ldots,x_d) = \prod_{k=1}^d f_k(x_k) \prod_{j=1}^{d-1} \prod_{i=1}^{d-j} c_{i,i+j | i+1,\ldots, i+j-1}(F(x_i|x_{i+1}, \ldots, x_{i+j-1}), F(x_{i+j}|x_{i+1}, \ldots, x_{i+j-1})) \end{equation} where the subscript $j$ identifies the tree, while $i$ runs over all edges in each tree. In contrast to this, in a D-vine, no node in any tree $T_j$ is connected to more than two edges yielding the decomposition \begin{equation}\label{decomposed-density d-vine} f(x) = \prod_{k=1}^d f_k(x_k) \prod_{j=1}^{d-1} \prod_{i=1}^{d-j} c_{j,j+i | 1,\ldots, j-1}(F(x_j|x_{1}, \ldots, x_{j-1}), F(x_{j+i}|x_{1}, \ldots, x_{j-1})). \end{equation} Examples of possible decompositions of a five-dimensional random vector via a C- and D-vine copula are shown in Figure \ref{fig:vine1} and \ref{fig:vine2}, respectively. \begin{center} --- insert Figures \ref{fig:vine1} and \ref{fig:vine2} here ---\\ \end{center} Fitting a vine copula model to a given dataset requires three separate steps. First, one needs to select the tree structure of the vine model. For a C- or D-vine, this amounts to the selection of a permutation of the indices $1,\ldots,d$ of the random variables. As such, for a $d$-dimensional random vector their exist $d!/2$ different C- and D-vines, respectively \citep[see][]{aas:2009a}.\footnote{Choosing the best fitting tree structure manually thus quickly becomes unfeasible in higher dimension \citep[see][]{dissmann:2011}.} Once a permutation has been chosen, the structure of the vine is fully specified. Second, the statistician has to select $d(d-1)/2$ bivariate pair-copulas from candidate copula families. In the last step, the parameters of the pair-copulas have to be estimated. To select the optimal tree structure, \citet{dissmann:2011} propose a heuristic procedure in which the tree structure is chosen via a maximum spanning tree algorithm that maximizes the sum of the absolute empirical Kendall's $\tau$ of all possible variable pairs on a given level of the tree. For the selection of the parametric pair-copulas, \citet{brechczado:2012} and \citet{dissmann:2011} propose a sequential heuristic which selects the best fitting parametric copula family for each pair-copula based on the candidate copulas' AIC values. Although this sequential selection of parametric pair-copulas using AIC does not necessarily yield a globally optimal AIC value for the vine, \citet{brechczado:2012} show that this heuristic yields considerably better results than a selection algorithm based on copula goodness-of-fit tests.\footnote{Results by \citet{weiss:2011} and \citet{grundke:2012} underline this finding. While \citet{weiss:2011} shows that goodness-of-fit tests give only little guidance for choosing copulas when one is interested in forecasting quantiles in the extreme tails of a joint distribution, the empirical results of \citet{grundke:2012} cast additional doubt on the ability of copula goodness-of-fit tests to identify stressed risk dependencies.} In the following, we concentrate on the problem of selecting the bivariate pair-copulas by substituting them with smooth nonparametric estimates of the underlying (pairwise) dependence structures. As a benchmark to our proposed nonparametric method, we employ the sequential heuristic by \citet{brechczado:2012} and \citet{dissmann:2011}. However, we expect our nonparametric approach to improve on the heuristic benchmark for several reasons. First, the nonparametric modeling of the pair-copulas eliminates completely the model risk of choosing an incorrect parametric family for a given pair-copula. Second, \citet{gronneberg:2012} prove that the use of AIC as a model selection criterion is not correct in case rank-transformed pseudo-observations are used (as is common in almost all applications of copulas in finance). Third, as already hinted at by \citet{dissmann:2011}, the incorrect specification of the parametric pair-copulas in the upper levels of a vine's tree structure can lead to a propagation and amplification of rounding errors causing the heuristic to become numerically unstable. In the next subsection, we define and discuss Bernstein copulas which we use as smooth nonparametric estimates of the pair-copulas in a vine model. \subsection{Bernstein copulas} As a nonparametric candidate for the pair-copulas in \eqref{decomposed-density c-vine} and \eqref{decomposed-density d-vine}, we consider the recently proposed Bernstein copulas. In the following, we briefly state some basic mathematical facts on Bernstein polynomials and Bernstein copulas, respectively. The Bernstein polynomials of degree $m$ are defined as \begin{equation} B(m, k, z) = \binom{m}{k} z^k (1-z)^{m-k}, \end{equation} where $k=0, \ldots, m \in \mathbb{N}$ and $0 \leq z \leq 1$. As our focus lies on the nonparametric modeling of pair-copulas in vines, we restrict our analysis in the following to bivariate Bernstein copulas. Let $U=(U_1, U_2)$ denote a discrete bivariate random vector with uniform margins over $T_i:=\{0, 1, \ldots\, m_i\}$ with grid size $m_i \in \mathbb{N}$ and $i=1,2$. In our analysis, we later choose $m_1:=m_2:=m=const$. With \begin{equation} p(k_1, k_2):= P(\displaystyle{\bigcap_{i=1}^2 \{U_i = k_i\}}), \quad (k_1, k_2) \in [0,1]^2 \end{equation} we can then define the Bernstein copula density as \begin{equation}\label{bernstein copula density} c(u_1, u_2):= \sum_{k_1=0}^{m_1-1} \sum_{k_2=0}^{m_2-1} p(k_1, k_2) \prod_{i=1}^2 m_i B(m_i-1, k_i, u_i), \quad (u_1, u_2) \in [0,1]^2. \end{equation} \citet{pfeifer:2009} show by integrating expression (\ref{bernstein copula density}) that the cdf of the Bernstein copula is then given by \begin{eqnarray}\label{bernstein copula} C(x_1, x_2) & := & \int_0^{x_2} \int_0^{x_1} c(u_1, u_2) d u_1 d u_2 \\ & = & \sum_{k_1=0}^{m_1} \sum_{k_2=0}^{m_2} P(\bigcap_{i=1}^2 \{U_i < k_i\}) \prod_{i=1}^2 B(m_i, k_i, u_i) \end{eqnarray} for $(x_1, x_2) \in [0,1]^2$. Note that in order to smoothly approximate the distribution or density of a copula in (\ref{bernstein copula density}) and (\ref{bernstein copula}), very high degrees for the Bernstein polynomials have to be chosen. The Bernstein copula as defined above can then be used to approximate the empirical copula process as defined, e.g., by \citet{deheuvels:1979}. To be precise, we approximate nonparametrically the joint distribution of $U=(U_1, U_2)$ by using a bivariate sample $\{(X_i, Y_i)\}^n_{i=1}$ of the underlying copula of size $n$. Let $X_{(k)}$ be the $k$th order statistic of the sample. The empirical copula of \citet{deheuvels:1979,deheuvels:1981} is then defined as \begin{equation}\label{eq:empcop} C_n (x, y) = \begin{cases} C_n \left(\frac{i-1}{n}, \frac{j-1}{n}\right), & \frac{i-1}{n} \le x < \frac{i}{n}, \frac{j-1}{n} \le y < \frac{j}{n}\\ 1, & x=y=1, \end{cases} \end{equation} where $C_n(x,y)=\frac{1}{n} \sum_{k=1}^n 1_{\{ X_j \le X_{(i)}, Y_j \le Y_{(i)} \}}$, $C_n(0,\frac{i}{n}) = C_n (\frac{j}{n}, 0) = 0$ and $C_n (0, \frac{i}{n}) = 0$ ($i,j=1,2,...,n$). To fit the Bernstein copula to the sample from the empirical copula, we first need to calculate the relative frequency of the observations in each target cell of a grid with given grid size $m$. The outcome of this is the contingency table $[a_{kl}]_{k,l=1,\ldots,m}$. Note, however, that the resulting marginals of the data $[a_{kl}]$ do not need to be uniformly distributed and that the resulting approximation could therefore not be a copula. To circumvent this problem, \citet{pfeifer:2009} propose to transform the contingency table $[a_{kl}]$ into a (possibly suboptimal) new contingency table $[x_{kl}]$ with uniform marginals via a Lagrange optimization approach yielding \begin{equation} x_{ij} = a_{ij} - \frac{a_{\cdot j}}{m} - \frac{a_{i \cdot}}{m} + \frac{2}{m^2} \qquad \text{for } i,j = 1,\ldots,m, \end{equation} where the index $\cdot$ denotes summation. Note that the quality of the Lagrange solution is reduced by an increasing number of the sample size $n$.\footnote{In unreported results, the optimization strategy of \citet{pfeifer:2009} proved to yield only suboptimal results.}\ We therefore chose to employ a different optimization strategy to correct for the non-uniform distribution of the marginals. Consequently, we calculate the approximation $[x_{kl}]$ to the contingency table $[a_{kl}]$ by solving the following optimization problem: \begin{eqnarray}\label{opt problem} & \displaystyle{\sum_{k=1}^m \sum_{l=1}^m} (x_{kl} - a_{kl})^2 \longrightarrow \text{min}\\ \text{subject to} & & \notag\\ & \displaystyle{\sum_{k=1}^m x_{kj} = \sum_{l=1}^m x_{il}} = \frac{1}{m} \qquad \text{and} \qquad x_{ij} \geq 0 \quad \text{for } i,j=1,\ldots,m. \end{eqnarray} To solve for the $[x_{kl}]$, we make use of the quadratic optimization algorithm of \citet{goldfarb:1982,goldfarb:1983}. In preliminary tests, the found solutions to this optimization problem yielded significantly lower quadratic errors than the procedure initially proposed by \citet{pfeifer:2009} thus confirming the need for a more refined optimization strategy. To use Bernstein copulas as pair-copulas both in our simulation study and the empirical application, we require efficient algorithms for simulating and evaluating the density and distribution of a given vine copula. To this end, we adapt the algorithms initially proposed by \citet{aas:2009a} by substituting the parametric h-hunctions (i.e., the partial derivatives of the copula densities) in these algorithms by the partial derivatives of the fitted bivariate Bernstein copulas. \section{Simulations} In this section, we illustrate the superiority of the smooth nonparametric vine model over the sequential heuristic of \citet{brech:2012} and \citet{dissmann:2011} for selecting the pair-copulas in a vine parametrically. In particular, we are interested in the error of the approximations to a pre-specified true copula using both our smooth nonparametric model as well as a vine model calibrated with parametric pair-copulas. The setup of our simulation study follows the procedure laid out in \citet{shen:2008}, but differs in that way that we also consider a (heuristically calibrated) parametric benchmark approximation to the true vine copula. As a measure for the approximation error, we compare the pre-specified true pair-copulas of the vine model with the parametric and nonparametric approximations and use the average squared error (ASE) of the cdfs of all bivariate pair-copulas each taken at $m_1\times m_2$ uniform grid points in $I^2$, i.e., \begin{equation} \label{ASE} ASE:=\frac{2}{d(d-1)}\frac{1}{m_1\cdot m_2}\sum_{i=1}^{d(d-1)/2}\sum_{j=1}^{m_1}\sum_{k=1}^{m_2} \left( \hat{C}_i\left(\frac{j}{m_1+1},\frac{k}{m_2+1}\right)-C_i\left(\frac{j}{m_1+1},\frac{k}{m_2+1}\right) \right)^2 \end{equation} where $C_i$ is the cdf of a pre-specified pair-copula from which we simulate a random sample of size $n$ and $\hat{C}_i$ is a (parametric or nonparametric) approximation to the pair-copula $C_i$ computed on the basis of the simulated random sample. In the simulation study, we consider two sample sizes $n=200$ and $n=500$ to assess the decreasing effect of the sample size on the approximation error. Furthermore, we analyze the effect of the type (C- or D-vine) as well as the dimensionality of the vine model on the approximation errors. To be precise, we simulate random samples from vines of dimension $d=3,5,6,7,11,13$. As the dimension of the vine model increases, so does the number of variables one has to condition on in the pair-copulas of the vine's lower trees. The pair-copulas in the lower trees of the vine, however, are generally more complex to model so that the accurate approximation of the pair-copulas on all levels of the vine constitutes a considerable challenge to our nonparametric approximation.\footnote{This is one reason why \citet{aas:2009a}, \citet{brech:2012} and \citet{dissmann:2011} propose to capture as much dependence of the joint distribution that is to be modeled in the first trees of a vine model. If these pair-copulas are modeled accurately, the remaining pair-copulas in the lower trees can then be truncated or simplified. Furthermore, the truncation and simplification of a vine on the lower levels of the vine's tree limits the potential propagation of rounding errors.}\ At the same time, the curse of dimensionality could additionally complicate the approximation of the pair-copulas thus making the comparison of our approximation for different dimensions a sensible exercise. Finally, we expect the propagation and amplification of rounding errors to increase in higher dimensions possibly leading to the numerical instability of the parametric heuristic. As candidate parametric copula families from which the pair-copulas of the true vine models are chosen, we use the Gaussian, Student's t, Clayton, Gumbel, Survival Clayton, Survival Gumbel, the rotated Clayton copula (90 degrees) and the rotated Gumbel copula (90 degrees). The pair-copulas as well as their respective parameters in each simulation are chosen randomly. For each sample size, dimension and vine type, we simulate $1,000$ random samples and approximate the data with a vine copula using Bernstein copulas as pair-copulas. As we are only interested in measuring the accuracy of the approximation of the pair-copulas, we calibrate our nonparametric vine using the correct vine type as well as the correct tree structure. As a benchmark, we calibrate a second vine copula by using the sequential heuristic proposed by \citet{brechczado:2012} and \citet{dissmann:2011}. Furthermore, we also compute the fraction of time the sequential procedure breaks down due to either the numerical instability of the evaluation of the likelihood function and the parameter estimation or due to the ASE tending to infinity. The results of the simulations are presented in Table \ref{tab:simu}. \begin{center} --- insert Table \ref{tab:simu} here ---\\ \end{center} The results shown in Table \ref{tab:simu} present several interesting insights into the finite sample properties of both the heuristically calibrated parametric and our proposed nonparametric vine copula models. First, we can see from Table \ref{tab:simu} that for lower dimensions (e.g., $d=3$ and $d=5$) the ASE of our nonparametric is considerably larger than for the parametric model calibrated by sequentially selecting the pair-copulas based on AIC values. With increasing dimension of the random vector, however, we can observe that the approximation error of the parametric model increases disproportionately compared to our proposed nonparametric model. Furthermore, the nonparametric model appears to be able to match the approximation error of the parametric approach for dimensions $d=13$ and higher. Most importantly, the parametric modeling approach becomes highly numerically unstable in higher dimensions. At the same time, our proposed nonparametric vine with Bernstein pair-copulas is extremely reliable yielding acceptable approximations to the true underlying dependence structure even for high-dimensional random vectors. The parametric approach, on the other hand, breaks down in approximately $50\%$ of all simulations for dimension $d=13$ and higher. In many of these cases, the bad approximation (or numerical instability) of the parametric approach was caused by the wrong selection of several parametric families for the pair-copulas in the vine model. Concerning the type of the vine copula model, we find no significant differences between the average approximation errors of the C- or D-vines. As expected, we also find the average approximation error of both the parametric and nonparametric model to be decreasing in the sample size used for estimating both models. To further illustrate the finding that the nonparametric model improves on the accuracy of a parametric vine especially in higher dimensions, we plot simulated samples from several parametrically and nonparametrically fitted copulas in Figure \ref{fig:approx} where we assume that the true underlying dependence structure is given by a Clayton copula with parameter $\theta=5$. From this copula, we simulate a random sample of size $n=500$ and fit both a nonparametric Bernstein copula as well as a parametric Clayton and Gumbel copula via Maximum-Likelihood to the data. From all three fitted copulas, we again simulate a random sample and compare the plots of the simulated observations with the original sample. \begin{center} --- insert Figure \ref{fig:approx} here ---\\ \end{center} The plots in Panels (a) and (b) in Figure \ref{fig:approx} show how the Bernstein copula approximately captures the lower tail dependence of the original sample shown in Panel (a). The plot of the Bernstein copula in Panel (b) also shows, however, that the nonparametric approximation of the data sample coincides with a loss in information on the tail behaviour of the true underlying dependence structure. At the same time, Panel (c) underlines the notion that the nonparametric model is not superior to a parametric model in which the parametric copula family has been chosen correctly. If, however, the parametric copula familiy is chosen incorrectly like it is shown in Panel (d), the wrong selection of the parametric copula family can cause considerable approximation errors and a severely inaccurate modeling of the underlying tail dependence. Given no prior information on the parametric copula familiy, the nonparametric Bernstein copula model clearly improves on the fit of an inaccurately fitted parametric model. The nonparametric modeling of the pair-copulas thus seems to be a sensible approach especially when the number of pair-copulas that need to be selected from candidate parametric copula families increases (i.e., with increasing dimension). \section{Empirical study} \subsection{Methodology} \label{sec:meth} The purpose of our empirical study is to investigate the superiority of the proposed smooth nonparametric vine copula models over the competing calibration strategy based on a sequential selection of parametric pair-copulas via AIC with regard to the accurate forecasting of a portfolio's VaR. The simulation study presented in the previous section has highlighted the finding that our vine model with smooth nonparametric pair-copulas is especially well-suited for dependence modeling in higher dimensions as the selection of parametric pair-copulas becomes numerically unstable due to error propagation and amplification. For low-dimensional problems, the heuristic selection of parametric copulas, however, seems to outperform our nonparametric approach with respect to the ASE of the vine copula's approximation. To show that our nonparametric model matches the results of the parametric heuristic even for low-dimensional problems, we concentrate in our empirical analysis on the VaR forecasts of a five-dimensional portfolio. Financial data are usually characterized by the presence of both conditional heteroscedasticity and asymmetric dependence in the log returns on financial returns. Therefore, we follow the vast majority of studies on copula models for VaR-estimation \citep{jondeau:2006,fantazzini:2009b,ausin:2010,hafner:2010} and employ standard GARCH(1,1)-models with Student's t-distributed innovations to model the marginal behaviour of our data. Although different specifications of the GARCH model are also possible, results found by \citet{hansen:2005} suggest that the choice of the order of a GARCH model is only of little importance for the model's forecasting accuracy. Throughout the empirical study, we consider continuous log returns on financial assets with prices $P_t$ $(t=0,1,\ldots,T)$. The assets' log returns $R_t$ are defined by $R_t:=\log(P_t/P_{t-1})$ for $t\geq 1$. Our focus lies on modeling the joint distribution of the $d$ assets, i.e., the joint distribution of the returns $R_{t1},\ldots,R_{td}$. The marginal behaviour of the assets is modeled by the use of GARCH(1,1)-models with t-distributed innovations. The marginal model is then given by \begin{eqnarray} R_{tj}&=&\mu_j+\sigma_{tj}Z_{tj},\\ \sigma^2_{tj}&=&\alpha_{0j}+\alpha_{1j}R^2_{t-1,j}+\beta_j\sigma^2_{t-1,j},\ j=1,\ldots,d;\ t=1,\ldots,T, \end{eqnarray} with independent and identically t-distributed innovations $Z_{tj}$. The dependence structure between the $d$ assets is introduced into the model by assuming the vector ${\bf Z}_t=(Z_{t1},\ldots,Z_{td})$ ($t=1,\ldots,T$) of the innovations to be jointly distibuted under a $d$-dimensional copula $C$ with \begin{equation} F_{\bf Z}({\bf z};\boldsymbol{\nu}_1,\ldots,\boldsymbol{\nu}_d,\boldsymbol{\omega}|\mathcal{F}_{t-1})=C\left[F_1(z_1;\boldsymbol{\nu}_1|\mathcal{F}_{t-1}),\ldots,F_d(z_d;\boldsymbol{\nu}_d|\mathcal{F}_{t-1});\boldsymbol{\omega}\right] \end{equation} where $\boldsymbol{\nu}_1,\ldots,\boldsymbol{\nu}_d$ are the parameter vectors of the innovations, $C$ is a copula and $\boldsymbol{\omega}$ is a vector of copula parameters (in case of the parametric model, otherwise $\boldsymbol{\omega}$ is simply empty). The parameters of the univariate GARCH-models are estimated via Quasi-Maximum Likelihood Estimation. For the estimation of both the nonparametric vine model as well as the parametric model calibrated by using the pair-copulas' AIC values, we make use of rank-transformed pseudo-observations rather than the original sample as input data.\footnote{For a comparative study on the finite sample properties of different ML-based estimators for copulas, see \citet{kim:2007}. The authors show that absent any information on the true distribution of the marginals, statistical inferences should be based on rank-transformed pseudo-observations.}\ As the main results for copulas only hold for i.i.d. samples, we use the parameter estimates for the univariate GARCH models and transform the original observations into standardized residuals to yield (approximately) i.i.d. observations before computing the pseudo-observations \citep{dias:2009}. In our empirical application, we consider an equally-weighted five-dimensional portfolio with returns $R_{p,t}=d^{-1}\sum_{j=1}^dR_{tj}$. The results from our simulation study underline the finding that our proposed vine copula model with Bernstein pair-copulas, on average, yields a better approximation to the empirical copula than the heuristically calibrated parametric model especially in higher dimensions. However, the parametric copula vine model could still outperform our proposed model w.r.t. the forecasting accuracy in low dimensions. In our empirical application, we therefore restrict our analysis to a portfolio consisting of five assets to additionally illustrate the nonparametric Bernstein vine copula model's superiority for low-dimensional problems. To forecast the portfolio returns, we employ the algorithm presented in the study by \citet{nikoloulopoulos:2011} initially proposed for in-sample forecasting which was extended to out-of-sample forecasting by \citet{weiss:2012rqfa}. The aim of the algorithm is the computation of a one-day-ahead forecast for the portfolio return $R_{p,t}$ via Monte Carlo simulation. In a first step, $K=10,000$ observations $u_{T+1,1}^{(k)},\ldots,u_{T+1,d}^{(k)}$ ($k=1,\ldots,K$) from the fitted (parametric or nonparametric) vine copula are simulated. Using the quantile function of the fitted marginal Student's t distributions, the simulated vine copula observations are then transformed into observations $z_{T+1,j}^{(k)}$ from the joint distribution of the innovations. In the next step, the simulated innovations are transformed into simulated returns $R_{T+1,j}^{(k)}=\hat{\mu}_j+\hat{\sigma}_{T+1,j}z_{T+1,j}^{(k)}$ where $\hat{\sigma}_{T+1,j}$ and $\hat{\mu}_j$ are the forecasted conditional volatility and mean values from the previously fitted marginal GARCH models. The MC-simulated forecasts of the portfolio return is then simply given by $R_{T+1,p}^{(k)}=d^{-1}\sum_{j=1}^dR_{T+1,j}^{(k)}$. Sorting the simulated portfolio returns for a given day in the forecasting period and taking the empirical one-day $\alpha$ percentile then yields the forecasted $\alpha\%$-VaR. To backtest the results of our forecasting, we employ the test of conditional coverage proposed by \citet{chris1} as well as two duration-based tests discussed in \citet{chris2}. \footnote{See \citet{Berk} for an excellent review of different methods for backtesting Value-at-Risk forecasts. A comparison of different backtests can be found in the recent study by \citet{escanciano}.} All three backtests are based on the hit sequence of VaR-exceedances which is defined by \[ h_{t,\alpha}:=\left\{ \begin{array}{ll} 1, & \mbox{if $R_{p,t}<$VaR$_{\alpha}(R_{p,t})|\mathcal{F}_{t-1}$} \\ 0, & \mbox{otherwise} \end{array}\right. . \] with $t$ being the time subscript and $\mathcal{F}_{t-1}$ being the set of available information. The test of conditional coverage by \citet{chris1} and \citet{chris2} jointly tests for the correct number of VaR-exceedances (unconditional coverage) and the serial independence of the violations over the complete out-of-sample (independence).\footnote{The test of unconditional coverage has been implicitly incorporated in the Basel Accord for determining capital requirements for market risks, see \citet{Basel}. Consequently, it has since become an industry standard in market risk management, see, e.g., \citet{escanciano}.} Under the null hypothesis of a correct number of VaR-exceedances that are independent over time, the hit sequence is simply distributed as \citep{chris2} \[ h_{t,\alpha} \sim i.i.d.\ Bernoulli(\alpha). \] Then, let $P$ be the length of the out-of-sample, $P_1$ be the number of VaR-exceedances and $P_0$ be the number of days on which the daily VaR forecast was not exceeded, respectively (and consequently $P=P_1+P_0$). The likelihood function for the i.i.d. $Bernoulli$ hit sequence with unknown parameter $\pi_1$ is \begin{equation} L(h_{t,\alpha},\pi_1)=\pi_1^{P_1}(1-\pi_1)^{P-P_1} \end{equation} and the Maximum-Likelihood estimate of $\pi_1$ is simply given by $\hat{\pi}_1=P_1/P$. The test of unconditional coverage is then given by a likelihood ratio test based on \begin{equation} LR_{UC}= -2 \left(\ln L(h_{t,\alpha},\hat{\pi}_1) - \ln L (h_{t,\alpha},\alpha)\right). \end{equation} To test the hypothesis of independently distributed hits, the hit sequence is assumed to follow a first order Markov sequence with switching probability matrix \begin{equation} \Pi = \begin{bmatrix} 1- \pi _{01} & \pi _{01}\\ 1- \pi _{11} & \pi _{11} \end{bmatrix} \end{equation} with $\pi_{ij}$ being the probability of an $i$ on day $t-1$ being followed by a $j$ on the next day $t$ and $i,j\in\left\{1;0\right\}$. Using the likelihood function \[ L (h_{t,\alpha},\pi_{01},\pi_{11}) = (1 - \pi_{01})^{P_0-P_{01}} \pi_{01}^{P_{01}} (1-\pi_{11})^{P_1-P_{11}}\pi_{11}^{P_{11}}. \] where $P_{ij}$ is the number of observations in $h_{t,\alpha}$ where a $j$ follows an $i$ and $i,j\in\left\{1;0\right\}$ and the ML-estimates $\hat{\pi}_{01}=P_{01}/P_0$ and $\hat{\pi}_{11}=P_{11}/P_1$, the likelihood ratio test of the independence of hits is given by \begin{equation} LR_{ind}= 2 \left(\ln L(h_{t,\alpha},\hat{\pi}_{01},\hat{\pi}_{11}) - \ln L (h_{t,\alpha},\hat{\pi}_{1})\right). \end{equation} Both tests are then combined via $LR_{CC}=LR_{UC}+LR_{ind}$ to yield the test of conditional coverage. We note here that we do not rely on the asymptotic Chi-squared distribution of the test statistic which is used, e.g., in the study by \citet{hsu:2011}. Although easy to implement, p-values derived under the assumption of the test statistic following a Chi-squared distribution are usually incorrect due to the generally low sample sizes when using hit sequences. Instead, we follow \citet{chris2} in generating approximate p-values via Monte Carlo-simulation. As an alternative to the test of conditional coverage, \citet{chris2} propose backtests based on the durations between VaR-exceedances. Then, let \begin{equation} D_i=t_i-t_{i-1} \end{equation} be the duration of time (in trading days) between two subsequent VaR-exceedances where $t_i$ is the time of the $i$th VaR-exceedance. Under the null hypothesis of a correctly specified VaR model, we would expect the process of no-hit durations to have no memory and mean $1/\alpha$. Consequently, the process $D$ of durations should follow an exponential distribution with $f_{exp}(D;\alpha)=\alpha\exp\left(-\alpha D\right)$.\footnote{See \citet{chris2} for details of the backtest and the motivation for using a continuous distribution for the discrete process $D$.} As an alternative hypothesis, \citet{chris2} propose to use the Weibull distribution for the process $D$ with $f_{W}(D;a,b)=a^bbD^{b-1}\exp\left(-(aD)^b\right)$ which nests the exponential distribution from the null hypothesis for $b=1$. Although this test potentially captures higher order dependence in the hit sequence $h_{t,\alpha}$, the information from the temporal ordering of the no-hit durations is not exploited in the backtest. As a remedy, \citet{chris2} propose a conditional duration-based test based on the Exponential Autoregressive Conditional Duration (EACD) model of \citet{engle:1998}. In the standard EACD (1,0) model, the conditional expected duration $\mathbb{E}_{i-1}(D_i)$ is assumed to follow the process \begin{equation} \mathbb{E}_{i-1}(D_i)\equiv\psi_i=\omega+\beta D_{i-1}. \end{equation} Again assuming an underlying exponential density with mean equal to one in the null hypothesis, the conditional distribution of the duration is is given by \begin{equation} f_{EACD}\left(D_i|\psi_i\right)=\frac{1}{\psi_i}\exp\left(-\frac{D_i}{\psi_i}\right). \end{equation} The null hypothesis of independent no-hit durations is then given by $H_0:\beta=0$. \subsection{Data} We consider a five-dimensional equal-weighted portfolio consisting of the returns on the EURO STOXX 50 Price Index, 30-year US Treasury Bonds, France Benchmark 10-year Government Bond Index, Gold Bullion LBM and one-month forward Crude Oil Brent. We obtain the data from the \textit{Thomson Reuters Datastream} database. We follow the screening procedure proposed by \citet{inceporter:2006} to control for known sources of data errors in \textit{Datastream}. To be precise, we check whether our data include prices below \$ 1 (which could lead to erroneous log returns due to \textit{Datastream}'s practice of rounding prices) as well as log returns above 300\% that are reversed within one month. Our five univariate time series do not suffer from any of these data errors. Our sample covers a period of $800$ trading days ranging from June 6, 2009 to July 19, 2012 and thus includes the aftermath of the default of Lehman Bros. as well as the onset of the Sovereign Debt Crisis. We use rolling windows with a length of $500$ trading days for forecasting the one-day-ahead VaR on the following trading day. Our full out-of-sample spans a period of $300$ trading days. Time series plots of the five portfolio constituents as well as the returns on the equal-weighted portfolio are shown in Figure \ref{fig:Hist1}. Panels (a) through (e) show the time series plots of the univariate returns, while Panel (f) shows the time series plot of the portfolio. The initial in-sample is shaded in grey to highlight the out-of-sample consisting of $300$ trading days. \begin{center} --- insert Figure \ref{fig:Hist1} here ---\\ \end{center} The plots in Figure \ref{fig:Hist1} show several distinct features that can complicate VaR-forecasting. First, all plots exhibit the common stylized fact of volatility clusters, e.g., in Panels (a) and (c). Second, overall volatility of the univariate returns differs significantly across our five portfolio constituents. For example, while the returns on the 30-year US treasury bonds are quite volatile in the in-sample and seem to calm in the out-sample, the opposite is true for the France Benchmark 10-year Government Bond Index which exhibits low volatility in the in-sample and a pronounced cluster of high volatility and extreme spikes around November 2011. This last result is, however, not surprising considering the fact that the Sovereign Debt Crisis experienced a climax at that time with the resignation of the Greek and Italian Prime Ministers, early elections in Spain and the expansion of the European Financial Stabilisation Mechanism (EFSM). Similarly, the price of Gold bullion became more volatile in the out-of-sample as well. The plot in Panel (f) shows that the combination of the five individual assets produces a portfolio which exhibits several phases of both high and low volatility as well as sudden extreme spikes in the portfolio's log returns. \subsection{Results} Following the methodology presented in Section \ref{sec:meth}, we compute the one-step-ahead forecasts of the portfolio-VaR for each day in the out-of-sample using rolling windows of $500$ trading days. To analyze the differential effect of different confidence levels for the VaR on our models' forecasting accuracy, we forecast the $2\%$-, $5\%$- and $10\%$-VaR for a long and the $97.5\%$-VaR for a short position in the portfolio. Thus, we would expect $6$, $15$, $30$ and $8$ exceedances below the forecasted VaRs, respectively.\footnote{For the short position, VaR-exceedances are defined as returns above the daily VaR forecast.}\ The VaR-forecasts as well as the realized portfolio returns for all four confidence levels are shown in Figures \ref{fig:forecasts1} and \ref{fig:forecasts2}. In both figures, Panels (a) and (b) show the realized portfolio returns and the VaR forecasts computed by the use of the nonparametric vine copula and the parametric vine copula model calibrated via the heuristic based on Akaike's Information Criterion, respectively. \begin{center} --- insert Figures \ref{fig:forecasts1} and \ref{fig:forecasts2} here ---\\ \end{center} The plots in Figures \ref{fig:forecasts1} and \ref{fig:forecasts2} show that both the nonparametric and the parametric model yield rather accurate forecasts of the portfolio's risk. While all VaR-forecasts are sufficiently close to the realized portfolio returns, exceedances of the VaR-forecasts occur only in case of large losses on the portfolio investment. Also, we can see that both the parametric and nonparametric model specifications yield quite similar VaR-forecasts. Thus, the proposed nonparametric vine copula model seems to perform exceptionally well even for low-dimensional portfolios. Furthermore, the finding of comparable VaR forecasts of both the nonparametric and parametric model remains valid for all four VaR confidence levels we consider. In addition to this, the results presented in Figures \ref{fig:forecasts1} and \ref{fig:forecasts2} also show that both models adequately adapt the VaR-forecasts to changes in the portfolio returns' volatility.\footnote{The flexible adjustment of both models to changes in return volatility also underlines the fact that a static dependence model in conjunction with dynamic marginal models suffices to model and forecast the dynamics of a multivariate return distribution.} To further assist in the interpretation of the results, Figures \ref{fig:exceed1} and \ref{fig:exceed2} highlight the positive and negative VaR-exceedances for all models and the three confidence levels for a long position in the portfolio. \begin{center} --- insert Figures \ref{fig:exceed1} and \ref{fig:exceed2} here ---\\ \end{center} Both figures underline our first impression from Figures \ref{fig:forecasts1} and \ref{fig:forecasts2} that both models forecast the VaR of the portfolio quite accurately. We can see from Figure \ref{fig:forecasts2} that not only do both models yield (approximately) correct numbers of VaR-exceedances for all three confidence levels, the exceedances also seem to occur randomly in time. Most importantly, however, our proposed nonparametric vine copula model with GARCH margins easily matches the heuristically calibrated parametric vine w.r.t. the accuracy of VaR-forecasting even for a relatively low-dimensional portfolio. To further substantiate this finding, we perform three formal backtests on the results of both the parametric and nonparametric vine models. The results of the three backtests are presented in Table \ref{tab:backtest}. \begin{center} --- insert Table \ref{tab:backtest} here ---\\ \end{center} The backtesting results stress our finding that both models yield comparable results. For example, all but one VaR models cannot be rejected at the $99\%$ confidence level based on the test of conditional coverage. Although the results of the unconditional duration-based backtest imply a significantly worse forecasting accuracy of both models, the p-values for both the nonparametric and the parametric model are comparable for different confidence levels of the VaR. This indicates that neither model outperforms the other one based on our second backtest. If we use the conditional duration-based test of \citet{christoffersenpelletier:2004} instead, none of the VaR-models is rejected. Turning to the number of VaR-exceedances, the results of our nonparametric vine copula model are slightly better for the $5\%$-VaR than those of the parametric benchmark while the opposite is true for the (for most practical uses too optimistic and thus unsuitable) $10\%$-VaR. Our backtesting results indicate that both models yield acceptable VaR-forecasts for a relatively low-dimensional portfolio. One could conclude from this finding that in general using our nonparametric vine copula model does not yield significantly better VaR-forecasts. However, one has to keep in mind that our empirical analysis was deliberately aimed at testing the hypothesis that the nonparametric model yields accurate VaR-forecasts even in lower dimensions. In unreported tests of high-dimensional portfolios, the parametric benchmark suffered from the same numerical instability that was also observed in our simulation study. At the same time, our nonparametric model produced accurate VaR-forecasts in a numerically stable fashion even for high-dimensional portfolios when the parametric benchmark had either broken down or produced woefully inaccurate VaR-forecasts. \section{Summary} In this paper, we propose to model the pair-copulas in a vine copula model nonparametrically by the use of Bernstein copulas. Our proposed model has the advantage of a significantly reduced model risk as it avoids the error-prone selection of pair-copulas from candidate parametric copula families. In contrast to previous studies on the use of discrete empirical copulas as pair-copulas, our proposed use of Bernstein copulas has the additional advantage that the building blocks in a vine model are approximated by smooth functions from which one can easily simulate random samples. We test the approximation error of the smooth nonparametric Bernstein vine copula model against a parametric benchmark calibrated by the use of a sequential heuristic based on AIC. The superiority of our proposed model is exemplified in an empirical risk management application. The results we find in our simulation study show that for low-dimensional problems, the parametric modeling approach outperforms our proposed nonparametric approach only marginally. However, the differences in the approximation error quickly vanish for higher dimensions with both models yielding comparable average approximations errors for dimensions $d=13$ and higher. At the same time, our proposed nonparametric vine copula model does not suffer from numerical instability and error propagation which plagues the parametric benchmark due to an increasing number of wrongly selected parametric pair-copulas. In the empirical risk management application, we test whether the differences in the average approximation error of the parametric and nonparametric vine copula models cause significant differences in both models' accuracy of forecasting the VaR of a low-dimensional asset portfolio. The results of our analysis show that even in lower dimensions ($d=5$), our nonparametric vine copula model yields VaR-forecasts that cannot be rejected by several different formal backtests. The proposed nonparametric vine copula model thus seems to match the (good) results of a parametric vine copula model in lower dimensions and significantly outperforms this benchmark in higher dimensions. A natural extension of our model would be to consider more sophisticated smooth approximations of the empirical copula. Cubic B-splines and non-uniform rational B-splines (NURBS) appear as natural candidates for this job. While Bernstein copulas have been shown to be good smooth nonparametric replacements for parametric pair-copulas, spline copulas should yield even better approximations while at the same time yielding numerically stable vine model calibrations as well. We intend to analyze the suitability of spline copulas in vine models in future research. \newpage \singlespacing \bibliographystyle{ecta}
\section{Introduction} Giant radio sources (GRSs) are defined as powerful extragalactic radio sources, hosted by galaxies or quasars, for which the projected linear size of their radio structure is larger than 0.72 Mpc{\footnote {Many authors, assuming $H_0=50$ km s$^{-1}$Mpc$^{-1}$, have used 1 Mpc as the defining size for GRSs. For the currently accepted cosmological parameters as given above, this size decreases to $\sim$0.72 Mpc.}} (assuming $H_0=71$ km s$^{-1}$Mpc$^{-1}$, $\Omega_M=0.27$, $\Omega_{\lambda}=0.73$; \citealt{b59}). Looking through the new, ``all-sky'' radio surveys such as the Westerbork Northern Sky Survey (\citealt{b51}), the NRAO VLA Sky Survey (NVSS; \citealt{b13}), the Faint Images of the Radio Sky at Twenty centimeters (FIRST; \citealt{b3}), the Sydney University Molonglo Sky Survey (\citealt{b10}) and the Seventh Cambridge Survey (\citealt{b41}) a large number of new giant sources were identified. Almost all of these GRSs are included in the samples of giants presented by \cite{b14}, \cite{b34}, \cite{b36}, \cite{b37}, \cite{b54}, \cite{b56}, as well as in the list of giants known before 2000 published by \cite{b25}. GRSs are very useful in studying a number of astrophysical problems, for example the evolution of radio sources, the properties of the intergalactic medium (IGM) at different redshifts, and the nature of the central active galactic nuclei (AGN). It is still unclear why such a small fraction of radio sources reach such a large size -- it may be due to special external conditions, such as lower IGM density, or due to the internal properties of the ``central engine''. Our knowledge about the nature of GRSs has improved somewhat following studies conducted in the last decade. However, these were focused almost exclusively on: the role of the properties of the IGM (\citealt{b1111, b32}), the advanced age of the radio structure (e.g. \citealt{b1112, b37b}), and recurrent radio activity (e.g. \citealt{b1114, b1130}) as responsible for gigantic size. To date, there are about 230 GRSs known, and just a small fraction of them ($\sim 8$ per cent) are actually related to quasars. The lobe-dominated radio quasars usually have a classical FRII (\citealt{b19}) morphology and most of their radio emission originates in the extended regions with steep radio spectra. The ratio of the flux density of the core to that of the lobes at 5 GHz is usually less than 1 (\citealt{b24}). Optically, the lobe-dominated radio quasars are similar to the core-dominated quasars (\citealt{b2}) and their lobe dominance is just an orientation effect. In general, it is believed that strong jet activity in an AGN is related to the parsec or sub-parsec scale condition of its host galaxy and specifically to the properties of its central black hole (BH; \citealt{b6, b8, b7}). Furthermore, if assuming that the power of the ``central engine'' is responsible for the linear size evolution of an radio source, we should expect the largest objects to be related to radio quasars, which host the most energetic AGNs, as opposed to radio galaxies. It is because there is observational evidence for a correlation between jet power and the expansion speed of a radio source's lobes (e.g. \citealt{b1119, b1120}). The jets of high-power sources carry a larger momentum flux, which in turn implies a greater flux of kinetic energy. Therefore, radio quasars, which are on average more luminous than radio galaxies, should have higher lobe expansion speeds than radio galaxies and, assuming similar mean lifetimes of both these types of AGNs, radio quasars have the potential to reach larger size. A typical lobe expansion speed could be of the order of a few hundredths (or more) of the speed of light (e.g. \citealt{b1121}), and this, together with the typical lifetime of an evolved radio source of the order of a few times $10^7$ years, gives a size of the order of a Mpc. The radio loudness of quasars still remains a debated issue. Radio observations of optically-selected samples of quasars showed that only 10-40 per cent of the objects are powerful radio sources (for reference see e.g. \citealt{b1122, b1123}). Recently, thanks to FIRST -- the large-area radio survey -- the number of quasars with faint radio fluxes has grown enormously. Therefore, it is now possible to investigate the optical and radio properties of quasars based on statistically large samples of objects (e.g. \citealt{b69, b1128, b1124, b1122, b1117, b1123}), and to try to understand the connection between the optical emission (luminosity, BH mass and spin, accretion rate) and the radio (jet) activity. Evidence that the spin of the BH plays a significant role in radio activity has recently been found (e.g. \citealt{b1115, b1116, b1117}). The relation between BH mass and radio loudness has also been intensively studied, but so far the results are equivocal. Many authors (e.g. \citealt{b35, b42, b17, b39, b44}) have found that, on average, radio-louder AGNs possess larger BH masses. However, there are also many reports arguing against any dependence between these quantities (e.g., \citealt{b48, b23, b71, b1122, b58}). Furthermore, the importance of the mechanical energy of jets and lobes released by BHs and the feedback on the surroundings has recently been realized (\citealt{b1125, b1127, b1126}). AGNs deposit large amounts of energy into their galactic environment which may, for example, be responsible for halting star formation. There is broad observational evidence that mechanical heating by jets plays an important role in balancing radiative losses from the intra-cluster medium. Radio jets and lobes of quasars can modify the structure of the environment not only on kpc scales but also, by the giant-size sources, on Mpc scales. The aim of this study is to investigate the radio and optical properties for a sample of lobe-dominated giant radio quasars (GRQs). We would like to answer the question whether the size of GRQs is related to the internal properties of their hosts. To investigate the role and importance of the central engine in generation of Mpc-scale structures we have compiled the largest sample of GRQs to date. The paper is organized as follows: in Sect.~2 we describe our source samples and in Sect.~3 the possible biases of the samples. In Sect.~4 we present definitions of the parameters used in the analysis; in Sect.~5 we investigate the relations between the optical and radio properties of our sources. Sect.~6 presents our conclusions.\\ Throughout the paper we assume the standard cosmology, with parameters as provided at the beginning of this Section. \section[]{The sample} In our analysis we use 45 GRQs of which 21 are taken from the existing literature (for details see Table 1). The remaining 24, which were not previously identified as GRQs, we selected from catalogues of radio quasars compiled by \cite{b20}, \cite{b4}, \cite{b69}, and \cite{b65}. The presented sample of giant-sized radio quasars is the largest to date. It contains sources even at large redshifts ($z\sim 2$). As a comparison sample, we selected 31 smaller lobe-dominated radio quasars from a list of radio sources given by \cite{b47}. In order to obtain a number of objects comparable to the GRQ sample, 18 quasars selected from the catalogues cited were added to the comparison sample. The linear sizes of these objects are close to the limiting size of 0.72 Mpc, as we wanted to have a smooth transition in linear size between the smaller radio quasars and the GRQs. The sources from the comparison sample of lobe-dominated radio quasar meet the following criteria: \begin{enumerate} \item Have optical spectra in the Sloan Digital Sky Survey (SDSS; \citealt{b1}). \item Possess the MgII(2798\AA) broad emission line in their spectra (as most of our GRQ spectra contain the MgII(2798\AA) line). This condition limits the range of the redshift to $0.4\lesssim z \lesssim 2$; it was adopted in order to have similar properties of the optical spectra for all quasars and hence allow homogeneous measurements using the same methods for both samples. \item Have a projected angular size of the radio structure larger than 0\farcm2, to properly separate the components (lobes and core) of the source in the FIRST maps (which have 5\arcsec$\times5$\arcsec\, angular resolution). \end{enumerate} \noindent All our quasars possess a classical FRII radio morphology. These objects lie almost in the plane of the sky and therefore the influence of relativistic beaming is small. It is easy to determine the physical size of such sources based on radio maps, even at high redshift. The final samples contain 45 GRQs and 49 smaller radio quasars, whose basic parameters are provided in Tables 1 and 2 respectively. The new, previously unrecognised, GRQs are marked in bold type in Table 1. Optical spectra from the SDSS as well as radio maps from the NVSS and FIRST surveys are available for almost all of these objects. In addition, the spectra of nine quasars published by \cite{b69}, \cite{b4}, \cite{b65} and \cite{b21} were provided to us in electronic FITS format by R. White (these are marked by the letter W in Tables 1 and 2). The columns of Table 1 and 2 contain respectively: (1) - J2000.0 IAU name; (2) and (3) J2000.0 right ascension and declination of the central position of the optical quasar; (4) redshift of the host object; (5) angular size in arcmin; (6) projected linear size in Mpc; (7) availability of the spectrum from the SDSS survey (S), or provided by White (W); availability of radio maps from NVSS or FIRST (N or F, respectively); (8) references to the identified object. Unfortunately, for two GRQs, J0631$-$5405 and J0810$-$6800, we had neither spectral nor radio data at hand, therefore we excluded them from further analysis. \begin{table*} \centering \begin{minipage}{130mm} \caption{List of giant-sized ($>$ 0.72 Mpc) radio quasars.} \begin{tabular}{@{}lccccccl@{}} \hline IAU & \hspace{0.5cm}$\alpha$(J2000.0)& $\delta$(J2000.0)& z & d & D & \hspace{0.5cm}Avail.& \hspace{0.5cm}Ref.\\ name & \hspace{0.5cm}(h m s) & ($^{o}$ \arcmin\, \arcsec) & & arcmin& Mpc & \hspace{0.5cm}Data & \\ (1) & \hspace{0.5cm}(2) & (3) &(4) &(5) & (6) & \hspace{0.5cm}(7) & \hspace{0.5cm}(8)\\ \hline {\bf J0204$-$0944} & \hspace{0.5cm}02 04 48.29 & $-$09 44 09.5 & 1.004 & 6.035 & 2.914 & \hspace{0.5cm}S,N,F& \hspace{0.5cm}1 \\ {\bf J0210$+$0118} & \hspace{0.5cm}02 10 08.26 & $+$01 18 42.3 & 0.870 & 2.618 & 1.214 & \hspace{0.5cm}W,N,F& \hspace{0.5cm}1 \\ J0313$-$0631 & \hspace{0.5cm}03 13 32.88 & $-$06 31 58.0 & 0.389 & 3.090 & 0.973 & \hspace{0.5cm}S,N & \hspace{0.5cm}2 \\ J0439$-$2422 & \hspace{0.5cm}04 39 09.20 & $-$24 22 08.0 & 0.840 & 1.960 & 0.899 & \hspace{0.5cm}N & \hspace{0.5cm}3 \\ J0631$-$5405 & \hspace{0.5cm}06 32 01.00 & $-$54 04 58.7 & 0.204 & 5.200 & 1.04 & \hspace{0.5cm}- & \hspace{0.5cm}4 \\ J0750$+$6541 & \hspace{0.5cm}07 50 34.43 & $+$65 41 25.4 & 0.747 & 3.271 & 1.439 & \hspace{0.5cm}N & \hspace{0.5cm}5 \\ {\bf J0754$+$3033} & \hspace{0.5cm}07 54 48.86 & $+$30 33 55.0 & 0.796 & 3.842 & 1.730 & \hspace{0.5cm}S,N,F& \hspace{0.5cm}6 \\ {\bf J0754$+$4316} & \hspace{0.5cm}07 54 07.96 & $+$43 16 10.6 & 0.347 & 8.061 & 2.360 & \hspace{0.5cm}S,N,F& \hspace{0.5cm}7 \\ {\bf J0801$+$4736} & \hspace{0.5cm}08 01 31.97 & $+$47 36 16.0 & 0.157 & 5.438 & 0.876 & \hspace{0.5cm}S,N,F& \hspace{0.5cm}7 \\ J0809$+$2912 & \hspace{0.5cm}08 09 06.22 & $+$29 12 35.6 & 1.481 & 2.184 & 1.118 & \hspace{0.5cm}S,N,F& \hspace{0.5cm}6, 8 \\ & & & & & & & \\ J0810$-$6800 & \hspace{0.5cm}08 10 55.10 & $-$68 00 07.7 & 0.231 & 6.500 & 1.42 & \hspace{0.5cm}- & \hspace{0.5cm}9 \\ J0812$+$3031 & \hspace{0.5cm}08 12 40.08 & $+$30 31 09.4 & 1.312 & 2.427 & 1.203 & \hspace{0.5cm}S,N,F& \hspace{0.5cm}8 \\ J0819$+$0549 & \hspace{0.5cm}08 19 41.12 & $+$05 49 42.7 & 1.701 & 1.923 & 0.987 & \hspace{0.5cm}S,N,F& \hspace{0.5cm}8 \\ J0842$+$2147 & \hspace{0.5cm}08 42 39.96 & $+$21 47 10.4 & 1.182 & 2.314 & 1.156 & \hspace{0.5cm}S,N,F& \hspace{0.5cm}8 \\ J0902$+$5707 & \hspace{0.5cm}09 02 07.20 & $+$57 07 37.9 & 1.595 & 1.678 & 0.862 & \hspace{0.5cm}S,N,F& \hspace{0.5cm}9, 8 \\ {\bf J0918$+$2325} & \hspace{0.5cm}09 18 58.15 & $+$23 25 55.4 & 0.688 & 2.079 & 0.885 & \hspace{0.5cm}S,N,F& \hspace{0.5cm}10\\ {\bf J0925$+$4004} & \hspace{0.5cm}09 25 54.72 & $+$40 04 14.2 & 0.471 & 4.379 & 1.546 & \hspace{0.5cm}S,N,F& \hspace{0.5cm}10\\ {\bf J0937$+$2937} & \hspace{0.5cm}09 37 04.04 & $+$29 37 04.8 & 0.451 & 2.640 & 0.909 & \hspace{0.5cm}S,N,F& \hspace{0.5cm}6, 10 \\ {\bf J0944$+$2331} & \hspace{0.5cm}09 44 18.80 & $+$23 31 18.5 & 0.987 & 1.870 & 0.899 & \hspace{0.5cm}S,N,F& \hspace{0.5cm}10\\ {\bf J0959$+$1216} & \hspace{0.5cm}09 59 34.49 & $+$12 16 31.6 & 1.089 & 1.964 & 0.966 & \hspace{0.5cm}S,N,F& \hspace{0.5cm}11\\ & & & & & & & \\ {\bf J1012$+$4229} & \hspace{0.5cm}10 12 44.29 & $+$42 29 57.0 & 0.364 & 3.088 & 0.933 & \hspace{0.5cm}S,N,F& \hspace{0.5cm}9 \\ {\bf J1020$+$0447} & \hspace{0.5cm}10 20 26.87 & $+$04 47 52.0 & 1.131 & 1.478 & 0.733 & \hspace{0.5cm}S,N,F& \hspace{0.5cm}11\\ {\bf J1020$+$3958} & \hspace{0.5cm}10 20 41.15 & $+$39 58 11.2 & 0.830 & 2.663 & 1.217 & \hspace{0.5cm}W,N,F& \hspace{0.5cm}9 \\ J1027$-$2312 & \hspace{0.5cm}10 27 54.91 & $-$23 12 02.0 & 0.309 & 2.860 & 0.774 & \hspace{0.5cm}N & \hspace{0.5cm}3 \\ J1030$+$5310 & \hspace{0.5cm}10 30 50.91 & $+$53 10 28.6 & 1.197 & 1.698 & 0.749 & \hspace{0.5cm}S,N,F& \hspace{0.5cm}8 \\ {\bf J1054$+$4152} & \hspace{0.5cm}10 54 03.27 & $+$41 52 57.6 & 1.090 & 4.702 & 2.314 & \hspace{0.5cm}S,N,F& \hspace{0.5cm}10\\ {\bf J1056$+$4100} & \hspace{0.5cm}10 56 36.26 & $+$41 00 41.3 & 1.785 & 1.543 & 0.791 & \hspace{0.5cm}S,N,F& \hspace{0.5cm}11\\ J1130$-$1320 & \hspace{0.5cm}11 30 19.90 & $-$13 20 50.0 & 0.634 & 4.812 & 1.977 & \hspace{0.5cm}N & \hspace{0.5cm}12\\ {\bf J1145$-$0033} & \hspace{0.5cm}11 45 53.67 & $-$00 33 04.6 & 2.054 & 2.642 & 1.340 & \hspace{0.5cm}S,N,F& \hspace{0.5cm}13\\ J1148$-$0403 & \hspace{0.5cm}11 48 55.89 & $-$04 04 09.6 & 0.341 & 3.265 & 0.945 & \hspace{0.5cm}N,F & \hspace{0.5cm}14\\ & & & & & & & \\ {\bf J1151$+$3355} & \hspace{0.5cm}11 51 39.68 & $+$33 55 41.8 & 0.851 & 2.083 & 0.959 & \hspace{0.5cm}S,N,F& \hspace{0.5cm}10\\ {\bf J1229$+$3555} & \hspace{0.5cm}12 29 25.56 & $+$35 55 32.5 & 0.828 & 1.672 & 0.761 & \hspace{0.5cm}S,N,F& \hspace{0.5cm}15\\ {\bf J1304$+$2454} & \hspace{0.5cm}13 04 51.42 & $+$24 54 45.9 & 0.605 & 2.431 & 0.977 & \hspace{0.5cm}W,N,F& \hspace{0.5cm}10\\ {\bf J1321$+$3741} & \hspace{0.5cm}13 21 06.42 & $+$37 41 54.0 & 1.135 & 1.531 & 0.759 & \hspace{0.5cm}S,N,F& \hspace{0.5cm}10\\ {\bf J1340$+$4232} & \hspace{0.5cm}13 40 34.70 & $+$42 32 32.2 & 1.343 & 2.309 & 1.173 & \hspace{0.5cm}S,N,F& \hspace{0.5cm}10\\ J1353$+$2631 & \hspace{0.5cm}13 53 35.92 & $+$26 31 47.5 & 0.310 & 2.803 & 0.761 & \hspace{0.5cm}W,N,F& \hspace{0.5cm}10, 16\\ {\bf J1408$+$3054} & \hspace{0.5cm}14 08 06.21 & $+$30 54 48.5 & 0.837 & 3.618 & 1.658 & \hspace{0.5cm}S,N,F& \hspace{0.5cm}10\\ {\bf J1410$+$2955} & \hspace{0.5cm}14 10 36.80 & $+$29 55 50.9 & 0.570 & 2.483 & 0.970 & \hspace{0.5cm}W,N,F& \hspace{0.5cm}6 \\ J1427$+$2632 & \hspace{0.5cm}14 27 35.61 & $+$26 32 14.5 & 0.363 & 3.822 & 1.158 & \hspace{0.5cm}S,N,F& \hspace{0.5cm}16\\ J1432$+$1548 & \hspace{0.5cm}14 32 15.54 & $+$15 48 22.4 & 1.005 & 2.824 & 1.364 & \hspace{0.5cm}S,N,F& \hspace{0.5cm}14\\ & & & & & & & \\ J1504$+$6856 & \hspace{0.5cm}15 04 12.77 & $+$68 56 12.8 & 0.318 & 3.140 & 0.867 & \hspace{0.5cm}N & \hspace{0.5cm}5 \\ J1723$+$3417 & \hspace{0.5cm}17 23 20.80 & $+$34 17 58.0 & 0.206 & 3.787 & 0.760 & \hspace{0.5cm}W,N,F& \hspace{0.5cm}17\\ J2042$+$7508 & \hspace{0.5cm}20 42 37.30 & $+$75 08 02.5 & 0.104 &10.052 & 1.138 & \hspace{0.5cm}N & \hspace{0.5cm}18\\ J2234$-$0224 & \hspace{0.5cm}22 34 58.76 & $-$02 24 18.9 & 0.550 & 3.236 & 1.241 & \hspace{0.5cm}N,F & \hspace{0.5cm}1 \\ {\bf J2344$-$0032} & \hspace{0.5cm}23 44 40.04 & $-$00 32 31.7 & 0.503 & 2.658 & 0.973 & \hspace{0.5cm}W,N,F& \hspace{0.5cm}1 \\ \hline \end{tabular}\\ References:(1) \cite{b4}; (2) \cite{b38}; (3) \cite{b25}; (4) \cite{b54}; (5) \cite{b34}; (6) \cite{b20}; (7) \cite{b55}; (8) \cite{b32}; (9) \cite{b65}; (10) \cite{b69}; (11) \cite{b31}; (12) \cite{b5}; (13) \cite{b33}; (14) \cite{b22}; (15) \cite{b57}; (16) \cite{b47}; (17) \cite{b26}; (18) \cite{b53}. \end{minipage} \end{table*} \begin{table*} \centering \begin{minipage}{125mm} \caption{List of smaller ($<$ 0.72 Mpc) radio quasars.} \begin{tabular}{@{}lccccccl@{}} \hline IAU & \hspace{0.5cm}$\alpha$(J2000.0) & $\delta$(J2000.0)& z & d & D &\hspace{0.5cm}Avail. & Ref.\\ name & \hspace{0.5cm}(h m s) & ($^{o}$ \arcmin\, \arcsec) & & arcmin & Mpc & \hspace{0.5cm}Data & \\ (1) & \hspace{0.5cm}(2) & (3) & (4) &(5) & (6) & \hspace{0.5cm}(7) & (8) \\ \hline J0022$-$0145 & \hspace{0.5cm}00 22 44.29 &$-$01 45 51.1 & 0.691 & 1.432 & 0.610 &\hspace{0.5cm}N,F & 1\\ J0034$+$0118 & \hspace{0.5cm}00 34 19.18 &$+$01 18 35.8 & 0.841 & 1.364 & 0.664 &\hspace{0.5cm}W,N,F & 1\\ J0051$-$0902 & \hspace{0.5cm}00 51 15.12 &$-$09 02 08.5 & 1.265 & 1.379 & 0.696 &\hspace{0.5cm}S,N,F & 1\\ J0130$-$0135 & \hspace{0.5cm}01 30 43.00 &$-$01 35 08.2 & 1.160 & 1.306 & 0.650 &\hspace{0.5cm}W,N,F & 1 \\ J0245$+$0108 & \hspace{0.5cm}02 45 34.07 &$+$01 08 14.2 & 1.537 & 0.883 & 0.453 &\hspace{0.5cm}S,N,F & 3\\ J0745$+$3142 & \hspace{0.5cm} 07 45 41.66 &$+$31 42 56.5 & 0.461 & 1.795 & 0.626 &\hspace{0.5cm}S,N,F & 3\\ J0811$+$2845 & \hspace{0.5cm}08 11 36.90 &$+$28 45 03.6 & 1.890 & 0.507 & 0.259 &\hspace{0.5cm}S,N,F & 3\\ J0814$+$3237 & \hspace{0.5cm}08 14 09.23 &$+$32 37 31.7 & 0.844 & 0.239 & 0.187 &\hspace{0.5cm}S,N,F & 3\\ J0817$+$2237 & \hspace{0.5cm}08 17 35.07 &$+$22 37 18.0 & 0.982 & 0.395 & 0.190 &\hspace{0.5cm}S,N,F & 3\\ J0828$+$3935 & \hspace{0.5cm}08 28 06.85 &$+$39 35 40.3 & 0.761 & 1.077 & 0.477 &\hspace{0.5cm}S,N,F & 3\\ & & & & & & & \\ J0839$+$1921 & \hspace{0.5cm}08 39 06.95 &$+$19 21 48.9 & 1.691 & 0.523 & 0.269 &\hspace{0.5cm}S,N,F & 3\\ J0904$+$2819 & \hspace{0.5cm}09 04 29.63 &$+$28 19 32.8 & 1.121 & 0.379 & 0.188 &\hspace{0.5cm}S,N,F & 3\\ J0906$+$0832 & \hspace{0.5cm}09 06 49.81 &$+$08 32 58.8 & 1.617 & 1.307 & 0.671 &\hspace{0.5cm}S,N,F & 4\\ J0924$+$3547 & \hspace{0.5cm}09 24 25.03 &$+$35 47 12.8 & 1.342 & 1.345 & 0.683 &\hspace{0.5cm}S,N,F & 5\\ J0925$+$1444 & \hspace{0.5cm}09 25 07.26 &$+$14 44 25.9 & 0.896 & 0.665 & 0.311 &\hspace{0.5cm}S,N,F & 3\\ J0935$+$0204 & \hspace{0.5cm}09 35 18.51 &$+$02 04 19.0 & 0.649 & 1.200 & 0.498 &\hspace{0.5cm}S,N,F & 3\\ J0941$+$3853 & \hspace{0.5cm}09 41 04.17 &$+$38 53 49.1 & 0.616 & 0.853 & 0.346 &\hspace{0.5cm}S,N,F & 3\\ J0952$+$2352 & \hspace{0.5cm}09 52 06.36 &$+$23 52 43.2 & 0.970 & 1.466 & 0.702 &\hspace{0.5cm}S,N,F & 2\\ J1000$+$0005 & \hspace{0.5cm}10 00 17.65 &$+$00 05 23.9 & 0.905 & 0.521 & 0.245 &\hspace{0.5cm}S,N,F & 3\\ J1004$+$2225 & \hspace{0.5cm}10 04 45.75 &$+$22 25 19.4 & 0.982 & 1.097 & 0.526 &\hspace{0.5cm}S,N,F & 3\\ & & & & & & & \\ J1005$+$5019 & \hspace{0.5cm}10 05 07.10 &$+$50 19 31.5 & 2.023 & 1.300 & 0.660 &\hspace{0.5cm}S,N,F & 2\\ J1006$+$3236 & \hspace{0.5cm}10 06 07.58 &$+$32 36 27.9 & 1.026 & 0.246 & 0.119 &\hspace{0.5cm}S,N,F & 3\\ J1009$+$0529 & \hspace{0.5cm}10 09 43.56 &$+$05 29 53.9 & 0.942 & 1.377 & 0.654 &\hspace{0.5cm}S,N,F & 2\\ J1010$+$4132 & \hspace{0.5cm}10 10 27.50 &$+$41 32 39.0 & 0.612 & 0.525 & 0.212 &\hspace{0.5cm}S,N,F & 3\\ J1023$+$6357 & \hspace{0.5cm}10 23 14.61 &$+$63 57 09.3 & 1.194 & 1.294 & 0.648 &\hspace{0.5cm}S,N,F & 6\\ J1100$+$1046 & \hspace{0.5cm}11 00 47.81 &$+$10 46 13.6 & 0.422 & 0.549 & 0.182 &\hspace{0.5cm}S,N,F & 3\\ J1100$+$2314 & \hspace{0.5cm}11 00 01.14 &$+$23 14 13.1 & 0.559 & 1.577 & 0.610 &\hspace{0.5cm}S,N,F & 5\\ J1107$+$0547 & \hspace{0.5cm}11 07 09.51 &$+$05 47 44.7 & 1.799 & 1.324 & 0.678 &\hspace{0.5cm}S,N,F & 2\\ J1107$+$1628 & \hspace{0.5cm}11 07 15.04 &$+$16 28 02.2 & 0.632 & 0.652 & 0.267 &\hspace{0.5cm}S,N,F & 3\\ J1110$+$0321 & \hspace{0.5cm}11 10 23.84 &$+$03 21 36.4 & 0.966 & 1.055 & 0.504 &\hspace{0.5cm}S,N,F & 3\\ & & & & & & & \\ J1118$+$3828 & \hspace{0.5cm}11 18 58.53 &$+$38 28 53.5 & 0.747 & 1.407 & 0.619 &\hspace{0.5cm}S,N,F & 5\\ J1119$+$3858 & \hspace{0.5cm}11 19 03.20 &$+$38 58 53.6 & 0.734 & 1.419 & 0.620 &\hspace{0.5cm}S,N,F & 5\\ J1158$+$6254 & \hspace{0.5cm}11 58 39.76 &$+$62 54 27.1 & 0.592 & 0.968 & 0.385 &\hspace{0.5cm}S,N,F & 3\\ J1217$+$1019 & \hspace{0.5cm}12 17 01.28 &$+$10 19 52.0 & 1.883 & 0.466 & 0.238 &\hspace{0.5cm}S,N,F & 3\\ J1223$+$3707 & \hspace{0.5cm}12 23 11.23 &$+$37 07 01.8 & 0.491 & 0.597 & 0.216 &\hspace{0.5cm}S,N,F & 3\\ J1236$+$1034 & \hspace{0.5cm}12 36 04.52 &$+$10 34 49.2 & 0.667 & 1.694 & 0.711 &\hspace{0.5cm}S,N,F & 3\\ J1256$+$1008 & \hspace{0.5cm}12 56 07.66 &$+$10 08 53.5 & 0.824 & 0.382 & 0.174 &\hspace{0.5cm}S,N,F & 3\\ J1319$+$5148 & \hspace{0.5cm}13 19 46.25 &$+$51 48 05.5 & 1.061 & 0.466 & 0.228 &\hspace{0.5cm}S,N,F & 3\\ J1334$+$5501 & \hspace{0.5cm}13 34 11.71 &$+$55 01 24.8 & 1.245 & 1.274 & 0.641 &\hspace{0.5cm}S,N,F & 3\\ J1358$+$5752 & \hspace{0.5cm}13 58 17.60 &$+$57 52 04.5 & 1.373 & 0.733 & 0.373 &\hspace{0.5cm}S,N,F & 3\\ & & & & & & & \\ J1425$+$2404 & \hspace{0.5cm}14 25 50.65 &$+$24 04 02.8 & 0.653 & 0.339 & 0.141 &\hspace{0.5cm}S,N,F & 3\\ J1433$+$3209 & \hspace{0.5cm}14 33 34.26 &$+$32 09 09.5 & 0.935 & 0.630 & 0.299 &\hspace{0.5cm}S,N,F & 3\\ J1513$+$1011 & \hspace{0.5cm}15 13 29.30 &$+$10 11 05.4 & 1.546 & 0.586 & 0.301 &\hspace{0.5cm}S,N,F & 3\\ J1550$+$3652 & \hspace{0.5cm}15 50 02.01 &$+$36 52 16.8 & 2.061 & 1.334 & 0.676 &\hspace{0.5cm}S,N,F & 4\\ J1557$+$0253 & \hspace{0.5cm}15 57 52.83 &$+$02 53 28.9 & 1.988 & 1.121 & 0.571 &\hspace{0.5cm}S,N,F & 2\\ J1557$+$3304 & \hspace{0.5cm}15 57 29.94 &$+$33 04 47.0 & 0.953 & 0.562 & 0.268 &\hspace{0.5cm}S,N,F & 3\\ J1622$+$3531 & \hspace{0.5cm}16 22 29.90 &$+$35 31 25.1 & 1.475 & 0.365 & 0.187 &\hspace{0.5cm}S,N,F & 3\\ J1623$+$3419 & \hspace{0.5cm}16 23 36.45 &$+$34 19 46.3 & 1.981 & 0.984 & 0.501 &\hspace{0.5cm}S,N,F & 2\\ J2335$-$0927 & \hspace{0.5cm}23 35 34.68 &$-$09 27 39.2 & 1.814 & 1.305 & 0.668 &\hspace{0.5cm}S,N,F & 1\\ \hline \end{tabular} References: (1) \cite{b4}; (2) \cite{b65}; (3) \cite{b47}; (4) \cite{b32}; (5) \cite{b69}; (6) \cite{b31}. \end{minipage} \end{table*} \section{Sample Biases} Due to the method used to complete our sample, the results may in some cases be influenced by selection effects, for example related to the sensitivity of the radio surveys used for selecting extended sources. The sample of giant radio quasars was compiled in three stages and each of them may be affected by bias. First, compact radio objects were selected, then the optical counterparts were checked for spectra typical of quasars. The selection criteria for these steps were described in detail in the papers referenced in Sect. 2 and we will not focus on them here. In the third stage of selection, we inspected the radio maps of several hundred candidates, looking for targets which have extended radio lobes in addition to radio cores. The NVSS and FIRST surveys have a completeness of 96 and 89 per cent and a reliability of 99 and 94 per cent to the $5\sigma$ limits of 2.3 and 1.0 mJy respectively (\citealt{b1131}). Since the resolution effect causes FIRST to become more incomplete for extended objects, we supplemented our search with the NVSS maps which have larger restoring beam size and hence larger surface brightness sensitivity. However, because of the limited baselines NVSS is insensitive to very extended coherent structures (larger than 15\arcmin). Fortunately, we do not expect the existence of objects with such large angular size, at least at high redshifts. In addition, extended and aged radio sources could have weak double lobes not connected with a visible bridge of high frequency radio emission. Therefore, it may be hard to recognise such a source as just one homogeneous object, especially at high redshift where the inverse Compton losses against the cosmic microwave background are large. Detecting a steep-spectrum and low surface-brightness radio bridge connecting the radio core with hot spots for distant objects is therefore challenging and this may have caused us to overlook some objects. It is worth noting that most of the recent works on quasars based on optical and radio data first select the candidates from optical catalogues of quasars and then correlate their coordinates with catalogues of radio sources. The authors usually concentrate on point-like radio sources, not extended objects (there are some exceptions, however, for example \citealt{b65}). \cite{b1123} considered extended radio structures, but analysed only those objects whose lobe separation was smaller than 1\arcmin. The authors stressed that the extended radio quasars represented a very small fraction of the SDSS quasars. They also wrote that quasars for which the radio structure diameter is greater then 1\arcmin are even rarer. Thus one has to realize that objects of the class studied here are extremely rare. As we pointed out in Sect.~2, the lobe-dominated radio quasars lie almost in the plane of the sky. Therefore, their measured radio luminosity is weakly influenced by relativistic beaming. In addition, it is easy to determine the proper physical size and volume occupied by the radio plasma for sources oriented in this manner. On the other hand, one should keep in mind that, besides lobe-dominated giant radio quasars as focused on here, there exist giant radio quasars located at a small angle to the line of sight which we have completely ignored because of the inability to determine their physical size. Given all the drawbacks described above, we have nonetheless shown that giant radio quasars do not comprise just a few objects as previously thought, but constitute a larger group. In addition to the sample of newly identified giants, we also added the set of previously known giant quasars to increase the number of objects tested. Summing up, the sample we presented here is limited by the described selection criteria and is not fully homogeneous. Therefore, applying the conclusions obtained here to the whole population of radio-loud quasars should be done with caution. \section[]{Data analysis} \subsection[]{Radio data} Using the Astronomical Image Processing System\footnote {http://www.aips.nrao.edu/} package for radio data reduction and analysis and maps from the NVSS and FIRST surveys, we measured the basic parameters of the selected radio quasars, which were further used to calculate their characteristics -- defined in the following way: \begin{enumerate} \item The arm-length-ratio, $Q$, which is the ratio of distances ($d_1$ and $d_2$) between the core and the hot spots (peaks of radio emission), normalized in such a way that always $Q>1$ (for details see Fig.~1). \item The bending angle, $B$, which is the complement of the angle between the lines connecting the lobes with the core. \item The lobes' flux-density ratio, $F=S_1/S_2$, where $S_1$ is the flux density of the lobe further from the core and $S_2$ is the flux density of the lobe closer to the core. \item The source total luminosity at 1.4~GHz, $P\rm_{tot}$, which is calculated following the formula given by \cite{b12}: \begin{eqnarray} \lefteqn{log P_{tot}(WHz^{-1})=log S_{tot}(mJy)-(1+\alpha) \cdot log(1+z)} \nonumber\\ & & {}+2log(D_L(Mpc))+17.08 \label{eq1} \end{eqnarray} \noindent where $\alpha$ is the spectral index (the convention we use here is $S_{\nu}\sim\nu^{\alpha}$) and $D_L$ is a luminosity distance. The total flux density, $S_{\rm tot}$, of individual sources is measured from NVSS maps and the average spectral index, in accordance with \cite{b68}, is taken for all sources as $\alpha=-0.6$. The core luminosity at 1.4~GHz, $P\rm_{core}$, is calculated in a similar manner, but instead of $S_{\rm tot}$ in equation~(\ref{eq1}) we substitute the core flux density, $S_{\rm core}$, which is measured from FIRST maps and the average spectral index value, according to \cite{b72}, is adopted as $\alpha = -0.3$. \item The inclination angle, $i$, which is the angle between the jet axis and the line of sight (i.e. $i = 90$\degr means that the object lies in the sky plane). The inclination angle was calculated, assuming that the Doppler boosting is the main factor underlying the asymmetries of a source, in the following way: \begin{equation} i=[acos(\frac{1}{\beta_j} \cdot \frac{(s-1)}{(s+1)})] \label{eq2} \end{equation} \noindent where $s=(S_j/S_{cj})^{1/2-\alpha}$, $S_j$ is the peak flux-density of the lobe closer to the core. $S_{cj}$ is the peak flux-density of the lobe further from the core and $\beta_j$ is the jet velocity. For all our objects, according to \cite{b68} and \cite{b1129}, we assume $\beta_j=0.6$c. \end{enumerate} \noindent The resulting values of the above parameters for our sources are listed in Table 3. For two objects, i.e. J0439$-$2422 and J1100$+$2314, we were not able to measure all the parameters, as for the source J0439$-$2422 the FIRST map was not available, and J1100$+$2314 has a too asymmetric radio structure. \begin{figure} \centering \includegraphics[width=0.75\linewidth, angle=0]{FIRST_c.eps} \caption{An example of a GRQ, J1321$+$3741. Radio contours are taken from the FIRST survey. Definitions of some parameters used for analysis are provided here (i.e. $B$, $S_1$, $S_2$, $d_1$, $d_2$) and also described in the text.} \end{figure} \subsection[]{Optical data} \subsubsection[]{Spectra reduction} The quasar spectra were reduced through the standard procedures of the Image Reduction and Analysis Facility \footnote {http://iraf.noao.edu/} package including galactic extinction and redshift correction. Each spectrum was corrected for galactic extinction taking into account values of the colour excess $E(B-V)$ and the $B$-band extinction, $A_B$, taken from the NASA/IPAC Extragalactic Database. We calculated the extinction parameter $R=E(B-V)/A_B$ for each quasar in our samples. The extinction-corrected spectrum was then transformed to its rest frame using the redshift value given in the SDSS (or if the SDSS spectrum was unavailable, from other publications). \subsubsection[]{Continuum subtraction and line parameters measurement} In order to obtain reliable measurements of emission lines, we need to subtract continuum emission, as optical and UV spectra of quasars are dominated by the power-law and Balmer continuum. Using the Image Reduction and Analysis Facility package, we subtracted the power-law continuum from our spectra. The continuum was fitted in several windows where we had not observed any emission lines (i.e. 1320--1350\AA, 1430--1460\AA, 1790--1830\AA, 3030--3090\AA, 3540--3600\AA\, and 5600--5800\AA). Particularly in the UV band, we also observe significant iron emission, which is often blended with the MgII(2798\AA) line. The procedure of subtracting the iron emission was similar to that described by \cite{b11}. We used an Fe template in the UV band (1250--3090\AA) as developed by \cite{b64}, and in the optical band (3535--7530\AA) given by \cite{b60}. First, we broadened the iron template by convolving it with Gaussian functions of various widths and multiplying by a scalar factor. Next, we chose the best fit of this modified template to each particular spectrum, and then subtracted it. After the subtraction of Fe line emission, we added the previously determined power-law continuum fit and refitted it once again (in a similar manner as suggested by \citealt {b64}). An example of a ``cleaned-up'' spectrum is presented in Fig.~2.\\ For the purpose of our analysis we needed to measure the parameters of broad emission lines like CIV(1549\AA), MgII(2798\AA) and H$_\beta$(4861\AA). In some cases, performing this measurement was difficult due to asymmetries in the line profiles (particularly of highly ionized lines such as CIV), where it was hard to fit a Gaussian function. In order to overcome the problem, we used the method described in \cite{b50}. In Tables 4 and 5 (cols.~2--4) we provide the respective widths of broad emission lines for GRQs and smaller quasars, respectively. We were unable to measure the MgII emission line parameters in the spectrum of the GRQ J1408+3054, as it showed strong broad-absorption features which considerably affected the emission line profile. \subsubsection[]{Black hole mass determination} The issue of determination of BH mass in AGNs has recently been often studied. The knowledge of the BH mass is of great importance in determining a number of physical parameters of AGNs and their evolution. In the first place, all of the commonly known techniques based on kinematic or dynamical studies (e.g. \citealt{b52}) are only useful for inactive galaxies. Therefore, they cannot be applied directly for AGNs, which are very luminous and distant. The most promising method for AGNs is reverberation mapping of the broad emission lines from the broad-line region (\citealt{b49}). This method works particularly well for type I AGNs (e.g. \citealt{b66}), where the broad line region is not obscured by a dusty and gaseous torus. Assuming that the gas in the broad-line region is virialized in the gravitational field of a BH, we can calculate its mass as: \begin{equation} M_{\rm BH}=\frac{R_{\rm BLR} V^2_{\rm BLR}}{\rm G} \label{eq3} \end{equation} \noindent where G is the gravitational constant, $R_{\rm BLR}$ is the distance from broad-line region clouds to the central BH, $V_{\rm BLR}$ is the broad-line region virial velocity, which can be estimated from the FWHM (Full Width at Half Maximum) of a respective emission line as: \begin{equation} V_{\rm BLR}=f \cdot FWHM \label{eq4} \end{equation} \noindent where $f$ is the scaling factor, which depends on structure, kinematics, and orientation of the broad line region (for randomly distributed broad line region clouds $f=\sqrt{3}/2$). Basing on this method, \cite{b28, b29} obtained an empirical relation between the broad line region size of an AGN and its optical continuum luminosity ($\lambda L_{\lambda}$) at 5100$\rm\AA$ (and later also at 1450$\rm\AA$, 1350$\rm\AA$ and in the 2--10~keV range): \begin{equation} R_{\rm BLR} \sim \lambda L_{\lambda}(5100\rm\AA)^{0.70 \pm 0.03} \label{aga5} \end{equation} \noindent This relation makes it possible to use an approximation to the reverberation mapping method, called the mass-scaling relation, which allows to determine BH mass using measurements of the FWHM of broad emission lines (e.g. CIV, MgII, H$\beta$) and the monochromatic continuum luminosity ($\lambda L_{\lambda}$) of a single-epoch spectrum only. In order to determine BH mass of our objects basing on the FWHM measurements of different emission lines, we applied the following equations: \begin{eqnarray} \lefteqn{M_{\rm BH}(CIV1549\rm\AA)= 4.57 \cdot 10^6 (\frac{\lambda L_{\lambda}(1350\rm\AA)}{10^{44}erg s^{-1}})^{0.53\pm0.06} \cdot} \nonumber\\ & & {}(\frac{FWHM(CIV1549\rm\AA)}{1000 km s^{-1}})^2 M_{\odot} \label{eq6} \end{eqnarray} \begin{eqnarray} \lefteqn{M_{\rm BH}(MgII2798\rm\AA)=7.24 \cdot 10^6(\frac{\lambda L_{\lambda}(3000\rm\AA)}{10^{44}erg s^{-1}})^{0.5} \cdot} \nonumber\\ & & {}(\frac{FWHM(MgII2798\rm\AA)}{1000 km s^{-1}})^2 M_{\odot} \label{eq7} \end{eqnarray} \begin{eqnarray} \lefteqn{M_{\rm BH}(H\beta4861\rm\AA)= 8.13 \cdot 10^6 (\frac{\lambda L_{\lambda}(5100\rm\AA)}{10^{44}erg s^{-1}})^{0.50\pm0.06} \cdot} \nonumber\\ & & {}(\frac{FWHM(H\beta4861\rm\AA)}{1000 km s^{-1}})^2 M_{\odot} \label{eq8} \end{eqnarray} \noindent Equations (\ref{eq6}) and (\ref{eq8}) were taken from \cite{b63}, while equation (\ref{eq7}) from \cite{b62}. The monochromatic continuum luminosities $\lambda L_{\lambda}$ can be computed as follows: \begin{equation} \lambda L_{\lambda} = 4 \pi D_{\rm Hubble}^2 \lambda f_{\lambda} \label{eq9} \end{equation} \noindent where $D_{\rm Hubble}$ is the comoving radial distance and $f_{\lambda}$ is the flux in the rest frame at wavelength $\lambda$ equal to 3000\AA, 5100\AA, or 1350\AA. The resulting rest frame fluxes, monochromatic continuum luminosities and BH masses for GRQs and smaller quasars are given in Table 4 and 5 (cols. 5--7, 8--10 and 11--13) respectively.\\ \begin{figure} \includegraphics[width=0.99\linewidth]{0809_5.eps} \caption{Spectrum of the giant radio quasar J0809+2912 and the best fit to the iron emission. The top spectrum is the observed spectrum in the rest frame overlaid with a power-law continuum, while the bottom one is the continuum-subtracted spectrum overlaid with the best fit to the iron emission.} \end{figure} \section[]{Results} \subsection[]{Radio properties} In our analysis we checked some general relations between radio parameters for our sample sources, similar to those shown for the sample of GRSs (mostly galaxies) described by \cite{b25}. On the optical- versus radio-luminosity plane our objects trace the regime of radio loudness (ratio of radio-to-optical luminosity) between 50 and 1000 and overlap with the FIRST-2dF sample of quasars of \cite{b1122}. In Fig.~3 we present the dependence between 1.4~GHz total luminosity and redshift for our quasars. It is important to note that our comparison sample of smaller radio quasars (sources marked as open circles in Fig.~3 and subsequent figures) contains only objects in the redshift range of 0.4$\lesssim$z$\lesssim$2 due to our selection criteria, i.e. the presence of the MgII(2798\AA) emission line in the spectra (for details see Sect.~2). Such a cut-off in the redshift range of the quasars from the comparison sample should not, however, affect our main results, since the majority of GRQs have redshifts in a similar range. Therefore, the non-existence of smaller radio quasars in the upper-left part of Fig.~3 is artificial, whereas the absence of GRQs in the lower-right corner of this figure is the result of sensitivity limit of the radio surveys which we used for source recognition and measurements of source's radio properties. It is known that in flux-limited samples we should expect a correlation between radio luminosity and redshift, since for larger distances we are able to detect only those sources which are luminous enough, and faint sources at higher redshifts are beyond the detection limit. For our quasar sample a dependence between redshift and total radio luminosity can be seen, but the correlation is not as strong as for the sample of GRSs from \cite{b25}. The Spearman rank correlation coefficient for the GRQs is 0.49, whereas for the GRSs from the paper cited above it is 0.90. This shows that the selection effects for our quasar sample are not as strong as for other radio galaxies and GRS samples of \cite{b25}, though they may still have affected some of our results. \begin{figure} \includegraphics[width=0.99\linewidth]{Ptotz_a.eps} \caption{1.4 GHz total radio luminosity as a function of redshift. The GRQs are marked with solid circles and quasars from the comparison sample are marked by open circles. J1623+3419, which is marked by a half-solid circle, has a projected linear size of 0.5 Mpc but, after correction for the inclination angle, its unprojected linear size is larger than the defining minimum size of GRSs.} \end{figure} In Fig.~4 we present the luminosity, $P$, versus linear size, $D$, relation. The $P$--$D$ diagram is a helpful tool in investigating the evolution of radio sources and was frequently used to test evolutionary models (e.g. \citealt{b27, b9}). In order to draw this diagram we used the unprojected linear size of the sources, which was derived by taking into account the inclination angle, $i$, as $D^* = D/sin(i)$, where $D$ is the projected linear size (given in Tables 1 and 2 derived as the sum of $d_1$ and $d_2$ - for details see Fig.~1). The diagrams show that GRQs have, on average, lower core and total radio luminosities. The trend which we observe in our $P$--$D$ diagrams is consistent with the predictions of evolutionary models and can suggest that, under favourable conditions, the luminous smaller, and probably younger, radio quasars may evolve in time into the lower-luminosity aged GRQs. The non-existence of objects in the bottom-left part of Fig.~4 may be due to selection effects. Because of the surface-brightness limit we may overlook some extended objects with low total radio luminosities. \begin{figure} \includegraphics[width=0.99\linewidth]{PtotLpr_1.eps}\\ \includegraphics[width=0.99\linewidth]{PcoreLpr_1.eps} \caption{Luminosity--linear size diagrams. The top panel shows the 1.4~GHz total radio luminosity and the bottom one shows core luminosity. The observed trend is consistent with predictions of evolutionary models.} \end{figure} In Fig.~5 we present the relation between the total and core radio luminosity. There is a strong correlation between those two quantities for radio quasars. We obtained a correlation coefficient of 0.76 and the slope of the linear fit equal to $0.82 \pm 0.08$, steeper than the slope of $0.59 \pm 0.05$ obtained by \cite{b25} for GRSs. The strong correlation between the core luminosity and the total luminosity in the population of giant-size radio galaxies was also mentioned by \cite{b37b}. On the one hand, this correlation can be attributed to the Doppler beaming of a parsec-scale jet and can reflect the different inclination angle of the nuclear jets, and thus inclination of the entire radio source's axis to the observer's line of sight. Relatively more luminous cores (in comparison to the total luminosity) should be observed in more strongly projected sources (i.e. quasars). Therefore, in GRQs one could expect to observe relatively stronger cores than in giant-sized radio galaxies. On the other hand, evolutionary effects (well visible in Fig.~4) can explain the clear difference in radio luminosity between GRQs and smaller quasars. Some authors (e.g. \citealt{b29}) have suggested that giants should have more prominent cores, as stronger nuclear activity is necessary to produce the larger linear sizes of their radio structure. \cite{b25} attempted to verify this hypothesis for giant-sized radio galaxies by plotting a diagram of the core prominence, $f_c$, which is the ratio of core luminosity to the total luminosity of the radio source, but found no trend of this kind. For GRQs investigated in this paper we also plotted such a diagram (see Fig.~6) and came to a similar conclusion. We can reconcile this with the existence of the core luminosity -- total luminosity correlation visible in Fig.~5 as a result of smaller quasars having more luminous cores but also larger total luminosities than GRQs. The resulting mean values of $f_c$ are 0.20 and 0.18 for GRQs and smaller quasars respectively. In Fig.~7 we plot the core prominence against linear size of the extended radio structure. The distribution of the core prominence is similar for GRQs and smaller radio quasars, which allows the claim that the strength of the central engine of GRQs is similar to that of smaller radio quasars. \begin{figure} \includegraphics[width=0.99\linewidth]{PcorePtot_1.eps} \caption{Core radio luminosity against the total radio luminosity for radio quasars. A strong correlation is visible. A linear fit to the data points is given by the line $logP_{\rm core}=(0.823\pm0.075)logP_{\rm tot}+(3.723\pm2.005)$.} \end{figure} \begin{figure} \includegraphics[width=0.99\linewidth]{fPtot_1.eps} \caption{Core prominence against total radio luminosity.} \end{figure} \begin{figure} \includegraphics[width=0.99\linewidth]{frac_1.eps} \caption{Core prominence against unprojected linear size. Also here, similar as in Fig~6, any correlation is visible.} \end{figure} We also investigated the asymmetries of radio structures in both our lobe-dominated radio quasar samples. It is well known that non-uniform environment (i.e. non-uniform density on both sides of the core) is one of the factors underlying radio structure asymmetries, which can be described by the arm-length ratio $Q$ (e.g. \citealt{bsch}). The distribution of this parameter for the GRQs and the quasars from the comparison sample is presented in Fig.~8. We found that the GRQs seem to be more symmetric than the smaller radio quasars (there were no GRQs with $Q > 2.4$ in our sample). However, the obtained mean values of the $Q$ parameter for GRQs and for the comparison sample are $1.41 \pm 0.33$ and $1.65 \pm 0.61$, respectively, and therefore indistinguishable within the error limits. This suggests that the IGM in which the giants evolve is not more symmetrical than that around the smaller sources. Our results for the large radio sources are comparable to those by \cite{b25}, which found that the mean value of the $Q$ parameter for the GRSs is 1.39, but for a comparison sample based on smaller 3CR sources they obtained a smaller $Q$ value equal to 1.19. We also compared the arm-length ratio value of GRQs and GRSs (see the bottom panel in Fig.~8). It can be seen clearly that the distributions for giant quasars and galaxies are similar. The values of the bending angle $B$ and lobe flux-density ratio $F$ give similar result for both samples of quasars with mean values of $B=7.40 \pm 5.89$, $F=1.45 \pm 1.16$ and $B=8.50 \pm 7.31$, $F=2.28 \pm 5.53$ for GRQs and comparison sample, respectively. In summary, there is no significant difference in the environmental properties of the IGM within which giant- and smaller-sized radio quasars evolve. Furthermore, we checked distribution of the inclination angle, $i$ (see Fig.~9). For our sample of radio quasars, we obtained that most objects have inclinations between 60$^{ \rm o}$ and 90$^{ \rm o}$. This result is inconsistent with the models of AGN unification scheme, where -- following \cite{b67} -- the inclination angle for quasars has a value between 0$^{ \rm o}$--45$^{ \rm o}$. In the objects with the angle larger than 45$^{ \rm o}$, the broad-line region should be partially or totally obscured by a dusty torus and the broad emission lines should not be as prominent as we observe in the spectra from our quasar sample. A plausible explanation of the observed distribution of inclinations is that there is no dusty torus in some AGNs (\citealt{b18}) or we are dealing with a clumpy, or receding torus (i.e. \citealt{b2222}), thus broad emission lines could have been observed even in quasars with large inclinations. The quasar with the largest asymmetry of its radio structure is J1623+3419 with $i=13^{ \rm o}$. Such a small value of the inclination angle can suggest that it should rather be classified as a BL~Lac object. Further observations are needed to confirm if its observed radio structure is actually related to a unique radio source. \begin{figure} \includegraphics[width=0.99\linewidth]{hist_popr_3resize_2.eps} \includegraphics[width=0.99\linewidth]{hist_popr_3bresize_2.eps} \caption{The distributions of the arm-length-ratio parameter $Q$. The top diagram shows all radio quasars from our samples, while the bottom one includes GRQs from our sample and GRSs taken from \citealt{b25}. The observed distribution of the Q parameter suggests that the IGM in which the giants evolve is not more symmetrical than that around the smaller sources.} \end{figure} \begin{figure} \includegraphics[width=0.99\linewidth]{inclination_2.eps} \caption{Distribution of the inclination angle, $i$, for the samples of radio quasars. For the definition of the inclination angle see Sect.~4.1. $i$=90\degr means that the jets and lobes lie in the plane of the sky.} \end{figure} \subsection[]{Black hole mass estimations} In order to obtain the central BH mass of quasars from our samples we used measurements of the CIV, MgII and H$\beta$ emission lines and the mass-scaling relations (equations (6), (7) and (8)). The mass values obtained are in the range of ~ $1.6 \cdot 10^8 M_{\odot} < M_{\rm BH} < 12.3 \cdot 10^8 M_{\odot}$ when using the MgII emission line, and $1.5 \cdot 10^8 M_{\odot} < M_{\rm BH} < 29.2 \cdot 10^8 M_{\odot}$ when using the $H_{\beta}$ emission line. For some GRQs and quasars from the comparison sample it was possible to compare the results obtained on the basis of different emission-line measurements. In Fig.~10 we present the relation between the mass values calculated from MgII vs $H_{\beta}$ lines and those from CIV vs MgII lines, respectively. We found that the mass estimations based on the MgII line on average tend to be smaller than those obtained using the $H_{\beta}$ emission line (the linear fit to the data points is given by the relation: $M_{BH}H_{\beta}=2.87(\pm 0.98)\cdot M_{BH}MgII+5.00(\pm 9.25)$), and the mass estimations based on CIV line are larger than those obtained from the MgII line ($M_{BH}CIV=0.68(\pm 0.14) \cdot M_{BH}MgII+1.08(\pm 0.61)$). The above results are consistent with the earlier comparisons of BH masses estimated by other authors (e.g. \citealt{b63, b16, b62}). \begin{figure} \centering \includegraphics[width=0.99\linewidth]{Mbhmghb_1.eps}\\ \hspace{0.5cm}\includegraphics[width=0.99\linewidth]{Mbhmgc4_1.eps} \caption{Comparison of BH mass values estimated using measurement of different emission lines. {\bf top diagram:} MgII versus $H_{\beta}$ BH masses; {\bf bottom diagram:} MgII versus CIV BH mass. The linear fits to data points are described in the text.} \end{figure} \subsection[]{Black hole mass vs radio properties} In the paper by \cite{bkp}, it is claimed that in the jet-formation models some dependence of jet power on BH mass should be expected. The assumption that the giants are formed due to a longer activity phase of the central AGN and/or more frequent duty cycles can imply that their BH masses should be larger because of longer accretion episodes. In Fig.~11 we present the relations between the total and core radio luminosity and the BH masse. It can be distinctly seen, however, that there is no correlation between the BH mass and either the core luminosity or the total luminosity for GRQs as well as smaller radio quasars. \begin{figure} \centering \includegraphics[width=0.99\linewidth]{MbhPcore_1.eps}\\ \hspace{1cm}\includegraphics[width=0.99\linewidth]{MbhPtot_1.eps} \caption{Relations between BH mass and radio luminosity at 1.4 GHz. {\bf top diagram:} BH mass vs core luminosity; {\bf bottom diagram:} BH mass vs total luminosity.} \end{figure} We also looked for a relation between BH masse and the unprojected radio linear size of a radio quasars. For the MgII BH mass estimations (Fig.~12), no obvious dependence has been found. Some interesting results can, however, be seen in Fig.~13. For the H$\beta$ and CIV BH mass estimations it can be clearly observed that the dependence between linear size of radio structures and their BH mass is quite significant. Surprisingly enough, the relation based on the H$\beta$ mass estimations for GRQs does not at all resemble that for quasars from the comparison sample. The slope of the linear fit for the sample of smaller quasars is steeper than that for the GRQs sample. This result suggests that the GRQs can be considered to represent another group of objects which differ physically from smaller quasars. We fitted linear functions independently to the data of GRQs and to the comparison sample (left panel of Fig.~13). The best fits obtained are as follows: $M_{\rm BH}H\beta = 10.995(\pm 7.023) \cdot D^*+1.629(\pm 10.268)$ and $M_{\rm BH}H\beta = 95.830(\pm 25.392) \cdot D^*-5.996(\pm 12.011)$ for the GRQs and for the comparison sample, with correlation coefficients of 0.48 and 0.74, respectively. We also plotted these lines on Fig.~12, taking into account the scaling factor between H$\beta$ and MgII BH mass estimations (equal to 2.87). It is obvious that the giants and the smaller radio quasars fulfil these relations quite well. Moreover, for the CIV mass estimation a weak correlation is also observed. The best fit is represented by a line $M_{\rm BH}CIV = 9.720(\pm 4.589) \cdot D^*+1.620(\pm 3.276)$ with a correlation coefficient of 0.51. The result obtained (particularly for the $H\beta$ mass estimations) can indicate that there may be some difference between GRQs and smaller radio quasars. It is hard to find a physical process to account for such a behaviour, especially as it is not found in the diagram for CIV BH masse. Some authors (e.g. \citealt{b30, b15}) suggested that the formation of different emission lines occurred in different regions of the broad line region, in the sense that the CIV emission should originate below the $H\beta$ emission. Therefore, GRQs and smaller radio quasars may differ with respect to the external structures of the broad line region, while their central parts would be similar. The question now is how to reconcile the fact that, according to the previously analysed relations for GRQs and smaller quasars, we did not see any clear distinction between these two types and here there is a clear difference. The possibility which comes to mind is that there is a difference in age between GRQs and smaller quasars and the composition of the broad-line region could be different for young and old quasars. However, the number of sources analysed using the CIV mass estimations is too small to allow for any definite conclusions, particularly relating to the smaller radio quasars. For example, this correlation deteriorates if we artificially shift the defining minimum GRQ size from 0.72 Mpc to a smaller value. Generally, apart from the above speculations on the composition of the broad line region, we can conclude that the apparent relationship between the linear size of the radio structure and the BH mass supports the evolutionary origin of GRQs: as time increases, the BH mass becomes larger and the size of radio structure grows. \begin{figure} \centering \includegraphics[width=0.99\linewidth]{MgLpr_1.eps} \caption{Dependence between the BH masses derived from the MgII emission line, and the unprojected linear sizes of the radio structures. The straight lines are reproduced from Fig.~13 (for details see the text).} \end{figure} \begin{figure} \hspace{0.3cm}\includegraphics[width=0.99\linewidth]{panel_1a.eps} \caption{Dependence between the BH mass derived from the H$\beta$ - {\bf left panel}, and the CIV - {\bf right panel} - emission line, and the unprojected linear size of the radio structure.} \end{figure} \subsection[]{Accretion rate} Using the obtained BH mass and the optical monochromatic continuum luminosity ($\lambda L_{\lambda}$) we calculated the accretion rate for our sample of quasars. The accretion rate is computed as \.m$(\lambda)$ = $L_{\rm bol}/L_{\rm Edd}$, where $L_{\rm bol}$ is the bolometric luminosity, assumed as: \begin{equation} L_{bol}=C_{\lambda} \lambda L_{\lambda}\label{eq10} \end{equation} \noindent where $C_{\lambda}$ is equal: 9.0 for ${\lambda}=5100\rm\AA$ (according to \citealt{b28}), 5.9 for ${\lambda}=3000\rm\AA$ (according to \citealt{b45}) and 4.6 for ${\lambda}=1350\rm\AA$ (according to \citealt{b61}). Following \cite{b16} the Eddington luminosity $L_{\rm Edd}$ is given by: \begin{equation} L_{\rm Edd}=1.45\cdot10^{38} M_{\rm BH}/M_{\odot} erg s^{-1} \label{eq11} \end{equation} The resulting values of $L_{\rm bol}$, $L_{\rm Edd}$ and \.m$(\lambda)$ for GRQs and smaller quasars are listed in Table 6 and 7, respectively. In Fig.~14 we present the BH mass as a function of accretion-rate values, which are calculated basing on the CIV, MgII and H$\beta$ emission lines as well as on the respective continuum luminosities, taking into account the scaling factor between H$\beta$, CIV and MgII mass estimations. As can be seen, the accretion rate is apparently higher for less massive BHs. A similar result was obtained by \cite{b16} for a sample of quasars and by \cite{b40} for narrow-line Seyfert galaxies. The result is consistent with the scenario of quasars increasing their BH mass during the accretion process solely. When there is no matter left, the accretion rate decreases, while a large amount of mass could have been accumulated in the central BH during the previous accretion episodes. In the scenario described by \cite{b40}, the accretion rate is high in the early stages of AGN evolution and drops later on, so we could expect that at higher redshifts we should observe objects with larger accretion rates. However, Fig.~15 shows that, for our samples of quasars, no dependence between accretion rate and redshift is seen.\\ The accretion rates for GRQs and for the comparison sample are consistent with typical values (0.01 $\div$ 1) for AGNs. Given the observed accretion rate we can constrain the lifetimes of the BHs in our samples. The obtained lower value for GRQs imply, that these sources are more evolved systems, for which the e-folding time to increase their BH mass (for a definition see e.g. \citealt{b2223}) is longer than in the case of smaller-size quasars. The obtained mean values of accretion rate (\.m(3000\AA)) are $0.07 \pm 0.03$ and $0.09 \pm 0.07$, respectively, for GRGs and smaller-size radio quasar. The dependence between accretion rate and unprojected linear size of radio structure is presented in Fig.~16. \begin{figure} \includegraphics[width=0.99\linewidth]{mmbh_poprawa1.eps} \caption{The dependence between BH mass and accretion rate \.m$(\lambda)$. The solid and open symbols mark GRQs and smaller-size radio quasar, respectively. Different symbols (circles, triangles and stars) represent estimations of accretion rate base on measurement of different emission lines (MgII, CIV and H$\beta$) and luminosities (at $\lambda=1350\rm\AA$, $\lambda=3000\rm\AA$ or $\lambda=5100\rm\AA$).} \end{figure} \begin{figure} \includegraphics[width=0.99\linewidth]{m3000z_2.eps} \caption{BH accretion rate vs redshift. No correlation is seen.} \end{figure} In Fig.~17 we present the dependence of accretion rate \.m(3000\AA) onto the core, as well as total, radio luminosity. There is a distinct trend for larger accretion rates to be observed in quasars with larger radio luminosity. The linear fits for \.m$(3000\rm\AA)$ are described by:\\ \.m(3000\AA)=0.114($\pm$0.044)log($P_{tot}$)-4.193($\pm$1.185),\\ \.m(3000\AA)=0.138($\pm$0.038)log$(P_{core})$-4.696($\pm$0.984)\\ with correlation coefficients equal to 0.29 and 0.39 respectively. \begin{figure} \centering \includegraphics[width=0.99\linewidth]{mLpr_2.eps}\\ \caption{Accretion rate vs unprojected linear size of radio structure.} \end{figure} \begin{figure} \includegraphics[width=0.99\linewidth]{m3000Ptot_1.eps}\\ \hspace{0.5cm}\includegraphics[width=0.99\linewidth]{m3000Pcore_1.eps} \caption{Accretion rate as a function of total radio luminosity - {\bf top panel} and core radio luminosity - {\bf bottom panel}. } \end{figure} \section[]{Conclusions} We have presented a comparison of radio and optical properties for a sample of GRQs and smaller radio quasars. It is important to mention that the measurements were obtained in a similar, homogeneous, manner for all sources from both the GRQ and comparison samples. Only the absolute values may be affected by some global calibration errors, if at all. The final conclusions are summarized below: \\ \begin{enumerate} \item Based on the $P$--$D$ diagram, we found that there is a continuous distribution of GRQs and smaller radio quasars. Therefore we can conclude that the GRQs could have evolved over time out of smaller radio quasars, which is consistent with the predictions of evolutionary models. We did not find that GRQs should have more prominent radio cores, which could suggest that the giants are similar to the smaller objects if we take into account their radio energetics.\\ \item The arm-length-ratio and bending angle values for both GRQs and smaller radio quasars are similar, which indicates that there is no significant difference of the environmental properties of the IGM within which giant- and smaller radio quasars evolve.\\ \item Statistically, the inclination angles obtained for our samples of quasars are inconsistent with traditional AGN unification scheme. Inclinations larger than 45$^{ \rm o}$ could, however, be explained based on recent results from studies of dusty torus properties.\\ \item The values of BH masses estimated here are similar to those for the powerful AGNs. The BH masses estimated using the MgII emission line are in the range of $1.6 \cdot 10^8 M_{\odot} < M_{\rm BH} < 12.2 \cdot 10^8 M_{\odot}$ and $1.0 \cdot 10^8 M_{\odot} < M_{\rm BH} < 20.3 \cdot 10^8 M_{\odot}$ for GRQs and for the smaller radio quasars respectively. We did not find any constraints for more massive BHs to be located in GRQs.\\ \item We did not find any significant correlation between the BH mass and the radio luminosity. However, using the $H_{\beta}$ and CIV line BH mass estimations a weak correlation between the linear size of the radio structure and the BH mass has been revealed. This might suggest that the linear size of giants could be related to their ``central engines''. Surprisingly enough, the same relation, but based on the $H_{\beta}$ analysis results, is different for the GRQs and for the smaller radio quasars, which could suggest an inherent difference between these types of objects. However, this result should be taken with some caution as it was obtained only for a small number of quasars. The relation between the linear size of the radio structure and the BH mass supports the evolutionary origin of GRQs.\\ \item The accretion rate for the more massive BHs is smaller than that for the less massive BHs. It is consistent with the scenario that quasars increase their BH mass during accretion process. The obtained mean value of accretion rate is equal to 0.07 for GRQs and 0.09 for smaller radio quasars. The lower value for GRQs suggests that GRQs are more evolved (aged) sources whose accretion process has slowed down or is almost over. The difference of \.m($\lambda$) and BH mass between the small-size radio quasars and large-size ones is, however, not significant, which could indicate similarities in their evolution. We found also a weak correlation between the accretion rates and the core radio luminosity, which confirms a connection between the accretion processes and the radio emission.\\ \item The results obtained from the measurements based on the H$\beta$ and CIV emission lines seem to be more homogeneous than those based on MgII. The BH masses derived from the H$\beta$ and CIV mass-scaling relations have smaller uncertainties than those of the MgII line. The large uncertainties in the case of MgII measurements are due to the fact that this line is strongly affected by the Fe emission. Moreover, the large uncertainty of the mass-scaling relation slope for the MgII line is also due to the absence of reverberation data from systematic monitoring.\\ \end{enumerate} In summary, taking into account the optical and radio properties, we can conclude that except for their size, the GRQs are similar to the smaller radio quasars. Their BH mass, accretion rate, prominence of radio core are comparable. The environment properties of GRQs and smaller radio quasars are also similar. Therefore, GRQs could be just evolved (aged) radio sources in which the accretion process has been diminished or is almost over and the large size is the consequence of their evolution. The sample of GRQs presented here, which is the largest one known to date, can be used for other astrophysical studies, such as on the evolution of radio sources. \section*{Acknowledgments} We are grateful to Richard White for providing us with a number of quasar spectra and Marianne Vestergaard for the template of the Fe emission. We thank J. Machalski, S. Zo{\l}a and D. Kozie{\l}-Wierzbowska for their detailed and very helpful comments on the manuscript. We thank also the anonymous referee for her/his very valuable comments. This project was supported in part by the Polish National Center of Science under decision DEC-2011/01/N/ST9/00726. \begin{table*} \centering \begin{minipage}{180mm} \caption{Parameters of radio structure for GRQs and smaller-size radio quasars from the comparison sample.} \begin{tabular}{lcccccc|lcccccc} \hline IAU &log(P$_{tot})$&log(P$_{core}$)& B&Q&F &i &IAU &log(P$_{tot})$ &log(P$_{core}$)&B&Q & F &i \\ name &W/Hz &W/Hz &[$^o$]& & &[$^o$]&name &W/Hz &W/Hz &[$^o$] & & & [$^o$]\\ (1) & (2) & (3) &(4) &(5) & (6) & (7) &(1) & (2) & (3) &(4) &(5) & (6) & (7)\\ \hline \multicolumn{7}{l}{Giant quasars} &\multicolumn{7}{l}{Comparison sample}\\ J0204$-$0944& 25.76& 24.92& 0.0 & 2.06& 0.59& 81 &J0022$-$0145& 26.62& 25.13 & 6.7 & 1.07& 2.08& 71\\ J0210$+$0118& 25.99& 25.31& 25.6& 1.38& 0.30& 63 &J0034$+$0118& 27.20& 24.51 & 7.9 & 1.90& 0.26& 60\\ J0313$-$0631& 25.89& 24.45& 5.6 & 1.11& 0.97& 87 &J0051$-$0902& 26.61& 24.60 & 9.1 & 1.39& 5.35& 55\\ J0439$-$2422& 27.09& $-$ & 4.5 & 1.67& 0.55& 79 &J0130$-$0135& 26.18& 24.68 & 5.9 & 1.90& 0.15& 62\\ J0750$+$6541& 26.39& 25.73& 5.5 & 1.05& 0.33& 65 &J0245$+$0108& 27.55& 25.85 & 12.0& 2.28& 0.68 & 85\\ J0754$+$3033& 26.10& 25.97& 18.8& 1.77& 1.20& 87 &J0745$+$3142& 26.96& 26.59 & 5.3 & 1.10& 0.62 & 88\\ J0754$+$4316& 25.63& 24.68& 0.2 & 1.07& 0.36& 85 &J0811$+$2845& 27.23& 26.71 & 6.1 & 2.35& 1.26 & 80\\ J0801$+$4736& 25.00& 24.58& 8.7 & 1.05& 2.21& 37 &J0814$+$3237& 26.84& 26.49 & 12.8& 2.40& 0.37 & 72\\ J0809$+$2912& 27.47& 26.21& 1.5 & 1.25& 0.04& 28 &J0817$+$2237& 27.69& 26.42 & 5.7 & 1.06& 2.38 & 71\\ J0812$+$3031& 26.07& 25.10& 2.4 & 1.37& 2.91& 71 &J0828$+$3935& 26.26& 24.94 & 0.4 & 1.15& 0.23 & 79\\ J0819$+$0549& 26.58& 25.19& 0.0 & 1.24& 1.54& 81 &J0839$+$1921& 27.78& 26.97 & 11.4& 1.36& 0.11 & 41\\ J0842$+$2147& 26.45& 25.46& 0.9 & 2.28& 0.92& 85 &J0904$+$2819& 26.83& 26.22 & 2.3 & 1.07& 3.15 & 45\\ J0902$+$5707& 26.53& 25.80& 9.3 & 1.38& 2.74& 79 &J0906$+$0832& 26.81& 26.05 & 0.5 & 1.29& 0.88 & 86\\ J0918$+$2325& 26.17& 25.59& 9.4 & 1.48& 0.99& 81 &J0924$+$3547& 26.49& 25.90 & 1.9 & 1.05& 0.68 & 84\\ J0925$+$4004& 25.73& 24.76& 5.7 & 1.13& 1.18& 80 &J0925$+$1444& 27.36& 26.04 & 5.6 & 1.26& 1.25 & 84\\ J0937$+$2937& 25.27& 24.13& 4.3 & 1.56& 0.69& 81 &J0935$+$0204& 27.06& 26.47 & 4.3 & 1.62& 37.20& -\\ J0944$+$2331& 26.95& 25.57& 7.1 & 1.67& 0.44& 83 &J0941$+$3853& 26.95& 26.03 & 0.3 & 1.44& 0.72 & 85\\ J0959$+$1216& 25.96& 25.21& 11.0& 1.17& 1.80& 73 &J0952$+$2352& 26.23& 26.02 & 17.7& 2.56& 0.97 & 89\\ J1012$+$4229& 25.52& 25.46& 15.6& 1.39& 3.71& 59 &J1000$+$0005& 27.46& 26.38 & 18.4& 1.35& 1.17 & 85\\ J1020$+$0447& 26.00& 27.85& 6.7 & 1.08& 3.71& 61 &J1004$+$2225& 27.40& 26.05 & 4.7 & 1.25& 1.03 & 85\\ J1020$+$3958& 25.39& 24.67& 1.9 & 1.12& 3.62& 58 &J1005$+$5019& 27.18& 26.96 & 16.2& 2.73& 1.54 & 76\\ J1027$-$2312& 26.21& 25.38& 8.2 & 1.15& 0.82& 88 &J1006$+$3236& 27.31& 26.97 & 0.9 & 2.89& 1.06 & 80\\ J1030$+$5310& 26.54& 25.70& 9.4 & 1.58& 3.22& 77 &J1009$+$0529& 26.79& 25.89 & 8.7 & 1.01& 1.75 & 78\\ J1054$+$4152& 25.82& 24.48& 18.8& 1.18& 4.23& 75 &J1010$+$4132& 27.35& 26.57 & 5.5 & 1.69& 12.73& 42\\ J1056$+$4100& 26.39& 25.63& 7.8 & 2.10& 0.89& 87 &J1023$+$6357& 27.01& 26.00 & 0.2 & 1.37& 0.44 & 70\\ J1130$-$1320& 27.20& 25.40& 0.7 & 1.13& 0.29& 67 &J1100$+$1046& 26.61& 26.19 & 1.8 & 1.20& 0.61 & 78\\ J1145$-$0033& 26.47& 25.72& 10.2& 1.29& 0.59& 82 &J1100$+$2314& 26.38& 25.14 & $-$ & $-$ & 3.64 & 68\\ J1148$-$0403& 26.32& 25.76& 12.8& 1.06& 1.20& 88 &J1107$+$0547& 26.42& 25.52 & 7.2 & 1.46& 0.63 & 80\\ J1151$+$3355& 26.26& 25.16& 11.3& 1.98& 0.33& 32 &J1107$+$1628& 27.09& 26.53 & 3.4 & 1.05& 0.67 & 87\\ J1229$+$3555& 26.20& 24.65& 14.9& 1.33& 0.39& 57 &J1110$+$0321& 27.35& 25.50 & 10.7& 2.38& 0.55 & 85\\ J1304$+$2454& 25.76& 25.41& 1.6 & 1.46& 1.82& 77 &J1118$+$3828& 26.10& 24.89 & 3.5 & 1.25& 3.25 & 55\\ J1321$+$3741& 26.53& 25.55& 17.0& 1.06& 0.72& 79 &J1119$+$3858& 26.43& 25.28 & 8.4 & 1.14& 2.66 & 64\\ J1340$+$4232& 26.24& 25.48& 3.5 & 1.92& 0.59& 89 &J1158$+$6254& 27.03& 25.32 & 4.3 & 1.53& 0.46 & 75\\ J1353$+$2631& 25.83& 24.78& 13.4& 1.21& 2.98& 34 &J1217$+$1019& 27.44& 26.13 & 26.8& 1.43& 0.73 & 87\\ J1408$+$3054& 25.93& 24.82& 4.9 & 1.41& 2.72& 80 &J1223$+$3707& 26.57& 25.48 & 3.9 & 1.75& 0.72 & 83\\ J1410$+$2955& 25.26& 24.56& 10.4& 1.30& 1.00& 81 &J1236$+$1034& 26.55& 25.18 & 0.4 & 1.42& 0.39 & 63\\ J1427$+$2632& 26.17& 25.23& 9.1 & 1.70& 0.40& 45 &J1256$+$1008& 26.98& 26.51 & 20.3& 1.14& 0.30 & 69\\ J1432$+$1548& 26.83& 25.71& 2.9 & 1.39& 0.99& 87 &J1319$+$5148& 27.70& 27.13 & 27.9& 1.81& 0.54 & 61\\ J1504$+$6856& 26.13& 25.52& 4.0 & 1.85& 1.66& 81 &J1334$+$5501& 27.46& 25.69 & 0.4 & 1.13& 0.79 & 89\\ J1723$+$3417& 26.26& 25.67& 1.1 & 1.05& 2.11& 51 &J1358$+$5752& 27.66& 25.86 & 1.3 & 1.20& 1.24 & 85 \\ J2042$+$7508& 25.67& 24.72& 7.1 & 1.03& 2.69& 61 &J1425$+$2404& 27.37& 26.62 & 18.9& 1.41& 1.60 & 83 \\ J2234$-$0224& 25.89& 24.71& 1.9 & 1.49& 0.18& 87 &J1433$+$3209& 26.89& 25.51 & 13.8& 1.14& 0.96 & 88 \\ J2344$-$0032& 25.46& 25.12& 0.8 & 1.54& 0.76& 79 &J1513$+$1011& 27.38& 26.36 & 20.9& 1.46& 1.07 & 83 \\ & & & & & & &J1550$+$3652& 26.99& 25.39 & 1.0 & 1.78& 0.28 & 64\\ & & & & & & &J1557$+$0253& 26.78& 26.40 & 7.4 & 3.62& 8.05 & 61\\ & & & & & & &J1557$+$3304& 27.45& 27.26 & 20.2& 1.54& 1.50 & 89 \\ & & & & & & &J1622$+$3531& 27.56& 26.36 & 17.2& 2.67& 0.59 & 82 \\ & & & & & & &J1623$+$3419& 26.31& 25.77 & 7.9 & 2.19& 0.64 & 13\\ & & & & & & &J2335$-$0927& 26.72& 26.11 & 8.0 & 2.74& 1.66 & 83\\ \hline \end{tabular} \end{minipage} \end{table*} \begin{table*} \centering \begin{minipage}{195mm} \caption{Parameters of optical spectra and black hole mass for GRQs.} \begin{tabular}{@{}lcccccccccccc@{}} \hline IAU & & FWHM & & &f$_\lambda$& & &Log$\lambda L_{\lambda}$& & &$M_{BH}$& \\ name & CIV & MgII & H$_\beta$& 1350\AA &3000\AA & 5100\AA &1350\AA & 3000\AA &5100\AA & CIV & MgII & H$\beta$ \\ & \multicolumn{3}{c|}{\rule[0mm]{4cm}{0.1pt}}& \multicolumn{3}{c|}{\rule[0mm]{3.5cm}{0.1pt}} & \multicolumn{3}{c|}{\rule[0mm]{3.9cm}{0.1pt}} & \multicolumn{3}{c|}{\rule[0mm]{4cm}{0.1pt}}\\ & \multicolumn{3}{c}{\AA} & \multicolumn{3}{c}{10$^{-17}$ erg cm$^{-2}$ s$^{-1}$ $\AA^{-1}$} & \multicolumn{3}{c|}{erg s$^{-1}$ } & \multicolumn{3}{c|}{10$^8$ M$\odot$} \\ \hline J0204$-$0944& $-$ & 34.19 & $-$ & $-$ &10.13 &\hspace{-0.5cm}$-$ & \hspace{-0.5cm}$-$& 44.61 & $-$ & \hspace{-0.5cm}$-$ & 1.93$^{\pm0.25}$ & $-$\\ J0210$+$0118& $-$ & 49.61 & $-$ & $-$ &46.57 &\hspace{-0.5cm}$-$ & \hspace{-0.5cm}$-$& 45.17 & $-$ & \hspace{-0.5cm}$-$ & 7.80$^{\pm0.22}$ & $-$\\ J0754$+$3033& $-$ & 48.59 & 139.90& $-$ &53.77 &\hspace{-0.5cm}10.65 & \hspace{-0.5cm}$-$& 45.17 & 44.70 & \hspace{-0.5cm}$-$ & 7.49$^{\pm0.25}$ & 13.56$^{\pm2.57}$\\ J0754$+$4316& $-$ & $-$ & 226.37& $-$ &$-$ &\hspace{-0.5cm}30.2 & \hspace{-0.5cm}$-$& $-$ & 44.53 & \hspace{-0.5cm}$-$ & $-$ & 29.17$^{\pm1.32}$\\ J0801$+$4736& $-$ & $-$ & 125.23& $-$ &$-$ &\hspace{-0.5cm}3.69 & \hspace{-0.5cm}$-$& $-$ & 42.97 & \hspace{-0.5cm}$-$ & $-$ & 1.48$^{\pm0.76}$ \\ J0809$+$2912& $-$ & 46.91 & $-$ & $-$ &44.45 &\hspace{-0.5cm}$-$ & \hspace{-0.5cm}$-$& 45.48 & $-$ & \hspace{-0.5cm}$-$ & 9.98$^{\pm0.51}$ & $-$\\ J0812$+$3031& $-$ & 30.91 & $-$ & $-$ &11.14 &\hspace{-0.5cm}$-$ & \hspace{-0.5cm}$-$& 44.72 & $-$ & \hspace{-0.5cm}$-$ & 1.80$^{\pm0.15}$ & $-$\\ J0819$+$0549& 75.57& 58.94 & $-$ & $-$ &1.50 &\hspace{-0.5cm}$-$ & \hspace{-0.5cm}$-$& 44.09 & $-$ & \hspace{-0.5cm}$-$ & 3.16$^{\pm2.77}$ & $-$\\ J0842$+$2147& $-$ & 33.16 & $-$ & $-$ &10.96 &\hspace{-0.5cm}$-$ & \hspace{-0.5cm}$-$& 44.74 & $-$ & \hspace{-0.5cm}$-$ & 2.12$^{\pm0.48}$ & $-$\\ J0902$+$5707& 31.09& 47.90 & $-$ & $-$ &10.40 &\hspace{-0.5cm}$-$ & \hspace{-0.5cm}$-$& 44.89 & $-$ & \hspace{-0.5cm}$-$ & 5.28$^{\pm1.28}$ & $-$\\ J0918$+$2325& $-$ & 55.91 & 173.43& $-$ &54.28 &\hspace{-0.5cm}8.92 & \hspace{-0.5cm}$-$& 45.08 & 44.52 & \hspace{-0.5cm}$-$ & 8.86$^{\pm0.46}$ & 16.95$^{\pm3.37}$\\ J0925$+$4004& $-$ & 62.30 & 196.95& $-$ &109.20 &\hspace{-0.5cm}20.40 & \hspace{-0.5cm}$-$& 45.10 & 44.60 & \hspace{-0.5cm}$-$ & 11.28$^{\pm4.24}$& 23.90$^{\pm2.01}$\\ J0937$+$2937& $-$ & 41.23 & 98.86 & $-$ &78.80 &\hspace{-0.5cm}14.70 & \hspace{-0.5cm}$-$& 44.92 & 44.42 & \hspace{-0.5cm}$-$ & 4.03$^{\pm0.51}$ & 4.91$^{\pm0.24}$ \\ J0944$+$2331& $-$ & 43.36 & $-$ & $-$ &34.92 &\hspace{-0.5cm}$-$ & \hspace{-0.5cm}$-$& 45.13 & $-$ & \hspace{-0.5cm}$-$ & 5.68$^{\pm0.54}$ & $-$\\ J0959$+$1216& $-$ & 46.95 & $-$ & $-$ &14.32 &\hspace{-0.5cm}$-$ & \hspace{-0.5cm}$-$& 44.81 & $-$ & \hspace{-0.5cm}$-$ & 4.59$^{\pm3.26}$ & $-$\\ J1020$+$0447& $-$ & 57.04 & $-$ & $-$ &6.56 &\hspace{-0.5cm}$-$ & \hspace{-0.5cm}$-$& 44.49 & $-$ & \hspace{-0.5cm}$-$ & 4.71$^{\pm1.45}$ & $-$\\ J1020$+$3958& $-$ & 66.70 & $-$ & $-$ &33.48 &\hspace{-0.5cm}$-$ & \hspace{-0.5cm}$-$& 45.00 & $-$ & \hspace{-0.5cm}$-$ & 11.52$^{\pm5.16}$& $-$\\ J1030$+$5310& $-$ & 36.46 & $-$ & $-$ &26.04 &\hspace{-0.5cm}$-$ & \hspace{-0.5cm}$-$& 45.13 & $-$ & \hspace{-0.5cm}$-$ & 3.99$^{\pm0.27}$ & $-$\\ J1054$+$4152& $-$ & 49.00 & $-$ & $-$ &23.03 &\hspace{-0.5cm}$-$ & \hspace{-0.5cm}$-$& 45.01 & $-$ & \hspace{-0.5cm}$-$ & 6.34$^{\pm3.40}$ & $-$\\ J1056$+$4100& 34.41& 51.61 & $-$ & $-$ &2.841 &\hspace{-0.5cm}$-$ & \hspace{-0.5cm}$-$& 44.39 & $-$ & \hspace{-0.5cm}$-$ & 3.43$^{\pm1.13}$ & $-$\\ J1145$-$0033& 61.12& $-$ & $-$ & 16.07&5.38 &\hspace{-0.5cm}$-$ & \hspace{-0.5cm}44.87 &44.74 & $-$ & \hspace{-0.5cm}18.42$^{\pm2.43}$& $-$ & $-$\\ J1151$+$3355& $-$ & 51.11 & $-$ & $-$ &29.15 &\hspace{-0.5cm}$-$ & \hspace{-0.5cm}$-$& 44.95 & $-$ & \hspace{-0.5cm}$-$ & 6.44$^{\pm2.48}$ & $-$\\ J1229$+$3555& $-$ & 31.93 & $-$ & $-$ &13.52 &\hspace{-0.5cm}$-$ & \hspace{-0.5cm}$-$& 44.60 & $-$ & \hspace{-0.5cm}$-$ & 1.67$^{\pm0.29}$ & $-$\\ J1304$+$2454& $-$ & 47.98 & 189.27& $-$ &79.83 &\hspace{-0.5cm}$-$ & \hspace{-0.5cm}$-$& 45.15 & $-$ & \hspace{-0.5cm}$-$ & 7.10$^{\pm0.43}$ & $-$\\ J1321$+$3741& $-$ & 73.96 & $-$ & $-$ &15.56 &\hspace{-0.5cm}$-$ & \hspace{-0.5cm}$-$& 44.87 & $-$ & \hspace{-0.5cm}$-$ & 12.23$^{\pm3.98}$& $-$\\ J1340$+$4232& $-$ & 52.94 & $-$ & $-$ &8.22 &\hspace{-0.5cm}$-$ & \hspace{-0.5cm}$-$& 44.69 & $-$ & \hspace{-0.5cm}$-$ & 5.12$^{\pm2.67}$ & $-$\\ J1353$+$2631& $-$ & 41.75 & 206.40& $-$ &136.10 &\hspace{-0.5cm}35.48 & \hspace{-0.5cm}$-$& 44.86 & 44.51 & \hspace{-0.5cm}$-$ & 3.87$^{\pm2.63}$ & 23.69$^{\pm2.59}$\\ J1410$+$2955& $-$ & 69.94 & $-$ & $-$ &47.28 &\hspace{-0.5cm}$-$ & \hspace{-0.5cm}$-$& 44.88 & $-$ & \hspace{-0.5cm}$-$ &11.04$^{\pm6.67}$& $-$\\ J1427$+$2632& $-$ & $-$ & 195.90& $-$ &$-$ &\hspace{-0.5cm}42.12 & \hspace{-0.5cm}$-$& $-$ & 44.72 & \hspace{-0.5cm}$-$ & $-$ & 27.08$^{\pm4.74}$\\ J1432$+$1548& $-$ & 52.83 & $-$ & $-$ &18.82 &\hspace{-0.5cm}$-$ & \hspace{-0.5cm}$-$& 44.88 & $-$ & \hspace{-0.5cm}$-$ & 6.34$^{\pm1.02}$ & $-$\\ J1723$+$3417& $-$ & $-$ & 64.88 & $-$ &168.40 &\hspace{-0.5cm}139.60& \hspace{-0.5cm}$-$& 44.62 & 44.77 & \hspace{-0.5cm}$-$ & $-$ & 3.16$^{\pm0.407}$ \\ J2344$-$0032& $-$ & 41.18 & $-$ & $-$ &70.72 &\hspace{-0.5cm}$-$ & \hspace{-0.5cm}$-$& 44.96 & $-$ & \hspace{-0.5cm}$-$ & 4.20$^{\pm0.31}$ & $-$\\ \hline \end{tabular} \end{minipage} \end{table*} \begin{table*} \centering \begin{minipage}{195mm} \caption{Parameters of optical spectra and black hole mass for smaller-size radio quasars.} \begin{tabular}{@{}lcccccccccccc@{}} \hline IAU & & FWHM & & &\hspace{-0.2cm}f$_\lambda$& & &Log$\lambda L_{\lambda}$& & &$M_{BH}$& \\ name & CIV & MgII & H$_\beta$ & \hspace{-0cm}1350\AA &\hspace{-0.2cm}3000\AA &\hspace{-0.4cm}5100\AA &1350\AA & 3000\AA &5100\AA & CIV & MgII & H$\beta$\\ & \multicolumn{3}{c|}{\rule[0mm]{4cm}{0.1pt}}& \multicolumn{3}{c|}{\rule[0mm]{3.5cm}{0.1pt}} &\multicolumn{3}{c|}{\rule[0mm]{3.9cm}{0.1pt}} & \multicolumn{3}{c|}{\rule[0mm]{4cm}{0.1pt}}\\ & \multicolumn{3}{c}{\AA} &\multicolumn{3}{c}{10$^{-17}$erg cm$^{-2}$s$^{-1}$$\AA^{-1}$} &\multicolumn{3}{c|}{erg s$^{-1}$ } & \multicolumn{3}{c|}{10$^8$ M$\odot$} \\ \hline J0034$+$0118& $-$ & 56.34& \hspace{-0.4cm}$-$& \hspace{-0.4cm}$-$& \hspace{-0.4cm}11.80 & \hspace{-0.05cm}$-$& \hspace{-0.5cm}$-$ &\hspace{-0.5cm}44.58 & \hspace{-0.5cm}$-$ & \hspace{-0.5cm}$-$ & \hspace{-0.05cm}5.06$^{\pm0.11}$ & $-$\\ J0051$-$0902& $-$ & 64.84& \hspace{-0.4cm}$-$& \hspace{-0.4cm}$-$& \hspace{-0.4cm}11.51 & \hspace{-0.05cm}$-$& \hspace{-0.5cm}$-$ & \hspace{-0.5cm}44.81 & \hspace{-0.5cm}$-$ & \hspace{-0.5cm}$-$ & \hspace{-0.05cm}8.73$^{\pm3.13}$ & $-$\\ J0130$-$0135& $-$ & 61.47& \hspace{-0.4cm}$-$& \hspace{-0.4cm}$-$& \hspace{-0.4cm}23.58 & \hspace{-0.05cm}$-$& \hspace{-0.5cm}$-$ & \hspace{-0.5cm}45.06 & \hspace{-0.5cm}$-$ & \hspace{-0.5cm}$-$ & \hspace{-0.05cm}10.56$^{\pm0.48}$ & $-$\\ J0245$+$0108& 35.54& 61.03& \hspace{-0.4cm}$-$& \hspace{-0.4cm}$-$& \hspace{-0.4cm}12.67 & \hspace{-0.05cm}$-$& \hspace{-0.5cm}$-$ & \hspace{-0.5cm}44.96 & \hspace{-0.5cm}$-$ & \hspace{-0.5cm}$-$ & \hspace{-0.05cm}9.23$^{\pm3.07}$ & $-$\\ J0745$+$3142& 186.40&53.84& \hspace{-0.4cm}173.80& \hspace{-0.4cm}$-$& \hspace{-0.4cm}499.50& \hspace{-0.05cm}107.50& \hspace{-0.5cm}$-$ & \hspace{-0.5cm}45.74 & \hspace{-0.5cm}45.30 & \hspace{-0.5cm}$-$ & \hspace{-0.05cm}17.67$^{\pm0.76}$& 41.92$^{\pm4.52}$\\ J0811$+$2845& 27.10& 54.05& \hspace{-0.4cm}$-$&\hspace{-0.4cm}59.34&\hspace{-0.4cm}12.92 & \hspace{-0.05cm}$-$&\hspace{-0.5cm}45.40 & \hspace{-0.5cm}45.08 & \hspace{-0.5cm}$-$ &\hspace{-0.5cm}6.88$^{\pm0.88}$&\hspace{-0.05cm}8.30$^{\pm1.16}$& $-$\\ J0814$+$3237& $-$ & 32.17& \hspace{-0.4cm}$-$& \hspace{-0.4cm}$-$& \hspace{-0.4cm}17.93 & \hspace{-0.05cm}$-$& \hspace{-0.5cm}$-$ &\hspace{-0.5cm}44.74 & \hspace{-0.5cm}$-$ & \hspace{-0.5cm}$-$ & \hspace{-0.05cm}1.99$^{\pm0.14}$ & $-$\\ J0817$+$2237& $-$ & 45.24& \hspace{-0.4cm}$-$& \hspace{-0.4cm}$-$& \hspace{-0.4cm}41.95 & \hspace{-0.05cm}$-$& \hspace{-0.5cm}$-$ &\hspace{-0.5cm}45.21 & \hspace{-0.5cm}$-$ & \hspace{-0.5cm}$-$ & \hspace{-0.05cm}6.72$^{\pm0.95}$ & $-$\\ J0828$+$3935& $-$ & 41.57& \hspace{-0.4cm}$-$& \hspace{-0.4cm}$-$& \hspace{-0.4cm}15.63 & \hspace{-0.05cm}$-$& \hspace{-0.5cm}$-$ &\hspace{-0.5cm}44.61 & \hspace{-0.5cm}$-$ & \hspace{-0.5cm}$-$ & \hspace{-0.05cm}2.85$^{\pm0.54}$ & $-$\\ J0839$+$1921& 21.60& 36.12& \hspace{-0.4cm}$-$& \hspace{-0.4cm}$-$& \hspace{-0.4cm}23.96 & \hspace{-0.05cm}$-$& \hspace{-0.5cm}$-$ &\hspace{-0.5cm}45.29 & \hspace{-0.5cm}$-$ & \hspace{-0.5cm}$-$ & \hspace{-0.05cm}4.72$^{\pm0.10}$ & $-$\\ J0904$+$2819& $-$ & 36.98& \hspace{-0.4cm}$-$& \hspace{-0.4cm}$-$& \hspace{-0.4cm}58.68 & \hspace{-0.05cm}$-$& \hspace{-0.5cm}$-$ &\hspace{-0.5cm}45.44 & \hspace{-0.5cm}$-$ & \hspace{-0.5cm}$-$ & \hspace{-0.05cm}5.88$^{\pm0.26}$ & $-$\\ J0906$+$0832& 38.41& 48.46& \hspace{-0.4cm}$-$& \hspace{-0.4cm}$-$& \hspace{-0.4cm}8.01 & \hspace{-0.05cm}$-$& \hspace{-0.5cm}$-$ &\hspace{-0.5cm}44.79 & \hspace{-0.5cm}$-$ & \hspace{-0.5cm}$-$ & \hspace{-0.05cm}4.78$^{\pm2.13}$ & $-$\\ J0924$+$3547& $-$ & 46.75& \hspace{-0.4cm}$-$& \hspace{-0.4cm}$-$& \hspace{-0.4cm}19.14 & \hspace{-0.05cm}$-$& \hspace{-0.5cm}$-$ &\hspace{-0.5cm}45.06 & \hspace{-0.5cm}$-$ & \hspace{-0.5cm}$-$ & \hspace{-0.05cm}6.09$^{\pm0.73}$ & $-$\\ J0925$+$1444& $-$ & 39.36& \hspace{-0.4cm}$-$& \hspace{-0.4cm}$-$& \hspace{-0.4cm}32.33 & \hspace{-0.05cm}$-$& \hspace{-0.5cm}$-$ &\hspace{-0.5cm}45.03 & \hspace{-0.5cm}$-$ & \hspace{-0.5cm}$-$ & \hspace{-0.05cm}4.18$^{\pm0.41}$ & $-$\\ J0935$+$0204& $-$ & 61.33& \hspace{-0.4cm}142.40& \hspace{-0.4cm}$-$& \hspace{-0.4cm}92.04 & \hspace{-0.05cm}17.1& \hspace{-0.5cm}$-$ &\hspace{-0.5cm}45.26 & \hspace{-0.5cm}44.76 & \hspace{-0.5cm}$-$ & \hspace{-0.05cm}13.22$^{\pm0.75}$ & 15.07$^{\pm1.44}$\\ J0941$+$3853& $-$ & 59.18& \hspace{-0.4cm}234.00& \hspace{-0.4cm}$-$& \hspace{-0.4cm}42.59 & \hspace{-0.05cm}9.19& \hspace{-0.5cm}$-$ &\hspace{-0.5cm}44.89 & \hspace{-0.5cm}44.45 & \hspace{-0.5cm}$-$ & \hspace{-0.05cm}8.01$^{\pm1.03}$ & 28.53$^{\pm2.11}$\\ J0952$+$2352& $-$ & 36.19& \hspace{-0.4cm}$-$& \hspace{-0.4cm}$-$& \hspace{-0.4cm}53.89 & \hspace{-0.05cm}$-$& \hspace{-0.5cm}$-$ &\hspace{-0.5cm}45.31 & \hspace{-0.5cm}$-$ & \hspace{-0.5cm}$-$ & \hspace{-0.05cm}4.85$^{\pm0.49}$ & $-$\\ J1000$+$0005& $-$ & 30.90& \hspace{-0.4cm}$-$& \hspace{-0.4cm}$-$& \hspace{-0.4cm}12.99 & \hspace{-0.05cm}$-$& \hspace{-0.5cm}$-$ &\hspace{-0.5cm}44.65 & \hspace{-0.5cm}$-$ & \hspace{-0.5cm}$-$ & \hspace{-0.05cm}1.65$^{\pm0.20}$ & $-$\\ J1004$+$2225& $-$ & 44.52& \hspace{-0.4cm}$-$& \hspace{-0.4cm}$-$& \hspace{-0.4cm}18.23 & \hspace{-0.05cm}$-$& \hspace{-0.5cm}$-$ &\hspace{-0.5cm}44.84 & \hspace{-0.5cm}$-$ & \hspace{-0.5cm}$-$ & \hspace{-0.05cm}4.29$^{\pm0.29}$ & $-$\\ J1005$+$5019& 18.18& 46.60& \hspace{-0.4cm}$-$&\hspace{-0.4cm}78.57&\hspace{-0.4cm}9.60 & \hspace{-0.05cm}$-$&\hspace{-0.5cm}$-$ & \hspace{-0.5cm}44.98 & \hspace{-0.5cm}$-$ &\hspace{-0.5cm}3.74$^{\pm0.46}$&\hspace{-0.05cm}5.53$^{\pm2.99}$& $-$\\ J1006$+$3236& $-$ & 34.69& \hspace{-0.4cm}$-$& \hspace{-0.4cm}$-$& \hspace{-0.4cm}8.954 & \hspace{-0.05cm}$-$& \hspace{-0.5cm}$-$ &\hspace{-0.5cm}44.57 & \hspace{-0.5cm}$-$ & \hspace{-0.5cm}$-$ & \hspace{-0.05cm}1.90$^{\pm0.20}$ & $-$\\ J1009$+$0529& $-$ & 68.01& \hspace{-0.4cm}$-$& \hspace{-0.4cm}$-$& \hspace{-0.4cm}71.05 & \hspace{-0.05cm}$-$& \hspace{-0.5cm}$-$ &\hspace{-0.5cm}45.41 & \hspace{-0.5cm}$-$ & \hspace{-0.5cm}$-$ & \hspace{-0.05cm}19.25$^{\pm1.07}$& $-$\\ J1010$+$4132& $-$ & 28.88& \hspace{-0.4cm}65.89& \hspace{-0.4cm}$-$& \hspace{-0.4cm}233.60& \hspace{-0.05cm}35.86& \hspace{-0.5cm}$-$ &\hspace{-0.5cm}45.62 & \hspace{-0.5cm}45.04 & \hspace{-0.5cm}$-$ & \hspace{-0.05cm}4.45$^{\pm0.34}$ & 4.45$^{\pm0.76}$\\ J1023$+$6357& $-$ & 48.48& \hspace{-0.4cm}$-$& \hspace{-0.4cm}$-$& \hspace{-0.4cm}53.00 & \hspace{-0.05cm}$-$& \hspace{-0.5cm}$-$ &\hspace{-0.5cm}45.43 & \hspace{-0.5cm}$-$ & \hspace{-0.5cm}$-$ & \hspace{-0.05cm}10.05$^{\pm0.97}$ & $-$\\ J1100$+$1046& $-$ & 49.35& \hspace{-0.4cm}21.30& \hspace{-0.4cm}$-$& \hspace{-0.4cm}60.94 & \hspace{-0.05cm}11.07& \hspace{-0.5cm}$-$ &\hspace{-0.5cm}44.76 & \hspace{-0.5cm}44.25 & \hspace{-0.5cm}$-$ & \hspace{-0.05cm}4.79$^{\pm2.17}$ & 0.19$^{\pm0.10}$\\ J1100$+$2314& $-$ & 66.57& \hspace{-0.4cm}296.03& \hspace{-0.4cm}$-$& \hspace{-0.4cm}100.50& \hspace{-0.05cm}31.02& \hspace{-0.5cm}$-$ &\hspace{-0.5cm}45.19 & \hspace{-0.5cm}44.91& \hspace{-0.5cm}$-$ & \hspace{-0.05cm}14.34$^{\pm4.02}$ & 77.29$^{\pm5.26}$\\ J1107$+$0547& 33.64& 40.26& \hspace{-0.4cm}$-$& \hspace{-0.4cm}$-$& \hspace{-0.4cm}6.66 & \hspace{-0.05cm}$-$& \hspace{-0.5cm}$-$ &\hspace{-0.5cm}44.76 & \hspace{-0.5cm}$-$ & \hspace{-0.5cm}$-$ & \hspace{-0.05cm}3.21$^{\pm1.52}$ & $-$\\ J1107$+$1628& $-$ & 36.98& \hspace{-0.4cm}88.07& \hspace{-0.4cm}$-$& \hspace{-0.4cm}180.00& \hspace{-0.05cm}33.76& \hspace{-0.5cm}$-$ &\hspace{-0.5cm}45.53 & \hspace{-0.5cm}45.04 & \hspace{-0.5cm}$-$ & \hspace{-0.05cm}6.57$^{\pm0.39}$ & 7.92$^{\pm0.86}$\\ J1110$+$0321& $-$ & 29.41& \hspace{-0.4cm}$-$& \hspace{-0.4cm}$-$& \hspace{-0.4cm}16.59 & \hspace{-0.05cm}$-$& \hspace{-0.5cm}$-$ &\hspace{-0.5cm}44.79 & \hspace{-0.5cm}$-$ & \hspace{-0.5cm}$-$ & \hspace{-0.05cm}1.77$^{\pm0.74}$ & $-$\\ J1118$+$3828& $-$ & 39.03& \hspace{-0.4cm}431.77& \hspace{-0.4cm}$-$& \hspace{-0.4cm}29.22 & \hspace{-0.05cm}3.46& \hspace{-0.5cm}$-$ &\hspace{-0.5cm}44.87 & \hspace{-0.5cm}44.17 & \hspace{-0.5cm}$-$ & \hspace{-0.05cm}3.39$^{\pm1.22}$ & 69.97$^{\pm29.70}$\\ J1119$+$3858& $-$ & 64.70& \hspace{-0.4cm}345.40& \hspace{-0.4cm}$-$& \hspace{-0.4cm}31.08 & \hspace{-0.05cm}6.27& \hspace{-0.5cm}$-$ &\hspace{-0.5cm}44.88 & \hspace{-0.5cm}44.41 & \hspace{-0.5cm}$-$ & \hspace{-0.05cm}9.47$^{\pm3.54}$ & 59.42$^{\pm20.60}$\\ J1158$+$6254& $-$ & 66.27& \hspace{-0.4cm}284.30& \hspace{-0.4cm}$-$& \hspace{-0.4cm}185.80& \hspace{-0.05cm}44.74& \hspace{-0.5cm}$-$ &\hspace{-0.5cm}45.50 & \hspace{-0.5cm}45.11 & \hspace{-0.5cm}$-$ & \hspace{-0.05cm}20.31$^{\pm2.35}$& 89.96$^{\pm19.39}$\\ J1217$+$1019& 20.56& 38.45& \hspace{-0.4cm}$-$&\hspace{-0.4cm}59.73&\hspace{-0.4cm}5.42 & \hspace{-0.05cm}$-$&\hspace{-0.5cm}45.39 & \hspace{-0.5cm}44.70 & \hspace{-0.5cm}$-$ &\hspace{-0.5cm}3.96$^{\pm0.53}$&\hspace{-0.05cm}2.71$^{\pm1.00}$& $-$\\ J1223$+$3707& $-$ & 50.07& \hspace{-0.4cm}264.00& \hspace{-0.4cm}$-$& \hspace{-0.4cm}34.43 & \hspace{-0.05cm}9.27& \hspace{-0.5cm}$-$ &\hspace{-0.5cm}44.63 & \hspace{-0.5cm}44.29 & \hspace{-0.5cm}$-$ & \hspace{-0.05cm}4.24$^{\pm0.95}$ & 30.00$^{\pm5.01}$\\ J1236$+$1034& $-$ & 38.83& \hspace{-0.4cm}272.90& \hspace{-0.4cm}$-$& \hspace{-0.4cm}53.47 & \hspace{-0.05cm}11.31& \hspace{-0.5cm}$-$ &\hspace{-0.5cm}45.05 & \hspace{-0.5cm}44.60 & \hspace{-0.5cm}$-$ & \hspace{-0.05cm}4.13$^{\pm1.94}$ & 46.03$^{\pm9.20}$\\ J1256$+$1008& $-$ & 33.46& \hspace{-0.4cm}33.74& \hspace{-0.4cm}$-$& \hspace{-0.4cm}15.93 & \hspace{-0.05cm}$-$& \hspace{-0.5cm}$-$ &\hspace{-0.5cm}44.67 & \hspace{-0.5cm}$-$ & \hspace{-0.5cm}$-$ & \hspace{-0.05cm}1.99$^{\pm0.18}$ & $-$\\ J1319$+$5148& $-$ & 39.67& \hspace{-0.4cm}$-$& \hspace{-0.4cm}$-$& \hspace{-0.4cm}76.29 & \hspace{-0.05cm}$-$& \hspace{-0.5cm}$-$ &\hspace{-0.5cm}45.52 & \hspace{-0.5cm}$-$ & \hspace{-0.5cm}$-$ & \hspace{-0.05cm}7.42$^{\pm0.80}$ & $-$\\ J1334$+$5501& $-$ & 55.70& \hspace{-0.4cm}$-$& \hspace{-0.4cm}$-$& \hspace{-0.4cm}21.12 & \hspace{-0.05cm}$-$& \hspace{-0.5cm}$-$ &\hspace{-0.5cm}45.06 & \hspace{-0.5cm}$-$ & \hspace{-0.5cm}$-$ & \hspace{-0.05cm}8.63$^{\pm0.92}$ & $-$\\ J1358$+$5752& $-$ & 45.54& \hspace{-0.4cm}$-$& \hspace{-0.4cm}$-$& \hspace{-0.4cm}67.78 & \hspace{-0.05cm}$-$& \hspace{-0.5cm}$-$ &\hspace{-0.5cm}45.62 & \hspace{-0.5cm}$-$ & \hspace{-0.5cm}$-$ & \hspace{-0.05cm}11.05$^{\pm1.92}$ & $-$\\ J1425$+$2404& $-$ & 56.62& \hspace{-0.4cm}131.60& \hspace{-0.4cm}$-$& \hspace{-0.4cm}89.40 & \hspace{-0.05cm}19.66& \hspace{-0.5cm}$-$ &\hspace{-0.5cm}45.25 & \hspace{-0.5cm}44.83 & \hspace{-0.5cm}$-$ & \hspace{-0.05cm}11.16$^{\pm1.62}$ & 13.87$^{\pm1.64}$\\ J1433$+$3209& $-$ & 45.42& \hspace{-0.4cm}$-$& \hspace{-0.4cm}$-$& \hspace{-0.4cm}2.924 & \hspace{-0.05cm}$-$& \hspace{-0.5cm}$-$ &\hspace{-0.5cm}44.02 & \hspace{-0.5cm}$-$ & \hspace{-0.5cm}$-$ & \hspace{-0.05cm}1.73$^{\pm1.85}$ & $-$\\ J1513$+$1011& 23.23& 42.70& \hspace{-0.4cm}$-$& \hspace{-0.4cm}$-$& \hspace{-0.4cm}36.67 & \hspace{-0.05cm}$-$& \hspace{-0.5cm}$-$ &\hspace{-0.5cm}45.42 & \hspace{-0.5cm}$-$ & \hspace{-0.5cm}$-$ & \hspace{-0.05cm}7.72$^{\pm0.57}$ & $-$\\ J1550$+$3652& 21.99& 47.58& \hspace{-0.4cm}$-$&\hspace{-0.4cm}41.78&\hspace{-0.4cm}4.80 & \hspace{-0.05cm}$-$&\hspace{-0.5cm}45.28 & \hspace{-0.5cm}44.69 & \hspace{-0.5cm}$-$ &\hspace{-0.5cm}3.97$^{\pm0.52}$&\hspace{-0.05cm}4.12$^{\pm1.81}$& $-$\\ J1557$+$0253& 10.42& 24.83& \hspace{-0.4cm}$-$&\hspace{-0.4cm}34.97&\hspace{-0.4cm}4.69 & \hspace{-0.05cm}$-$&\hspace{-0.5cm}45.19 & \hspace{-0.5cm}44.66 & \hspace{-0.5cm}$-$ &\hspace{-0.5cm}0.79$^{\pm0.16}$&\hspace{-0.05cm}1.09$^{\pm1.28}$& $-$\\ J1557$+$3304& $-$ & 52.94& \hspace{-0.4cm}$-$& \hspace{-0.4cm}$-$& \hspace{-0.4cm}25.46 & \hspace{-0.05cm}$-$& \hspace{-0.5cm}$-$ &\hspace{-0.5cm}44.97 & \hspace{-0.5cm}$-$ & \hspace{-0.5cm}$-$ & \hspace{-0.05cm}7.05$^{\pm1.15}$ & $-$\\ J1622$+$3531& $-$ & 43.02& \hspace{-0.4cm}$-$& \hspace{-0.4cm}$-$& \hspace{-0.4cm}10.56 & \hspace{-0.05cm}$-$& \hspace{-0.5cm}$-$ &\hspace{-0.5cm}44.86 & \hspace{-0.5cm}$-$ & \hspace{-0.5cm}$-$ & \hspace{-0.05cm}4.08$^{\pm0.83}$ & $-$\\ J1623$+$3419& 25.17& 48.20& \hspace{-0.4cm}$-$&\hspace{-0.4cm}14.37&\hspace{-0.4cm}2.59 & \hspace{-0.05cm}$-$&\hspace{-0.5cm}44.80 & \hspace{-0.5cm}44.40 & \hspace{-0.5cm}$-$ &\hspace{-0.5cm}2.88$^{\pm0.59}$&\hspace{-0.05cm}3.04$^{\pm1.78}$& $-$\\ J2335$-$0927& 14.18& 35.59& \hspace{-0.4cm}$-$&\hspace{-0.4cm}60.11&\hspace{-0.4cm}1.14 & \hspace{-0.05cm}$-$&\hspace{-0.5cm}45.38 & \hspace{-0.5cm}44.00 & \hspace{-0.5cm}$-$ &\hspace{-0.5cm}1.85$^{\pm0.26}$&\hspace{-0.05cm}1.04$^{\pm0.99}$& $-$\\ \hline \end{tabular} \end{minipage} \end{table*} \begin{table*} \centering \begin{minipage}{140mm} \caption{Optical luminosity and accretion rate for GRQs.} \begin{tabular}{@{}lccccccccc@{}} \hline IAU & &log(L$_{bol})$ & & & log(L$_{Edd}$) & &\.m1350 &\.m3000 & \.m5100\\ name &1350\AA &3000\AA & 5100\AA & CIV & MgII & $H\beta$ & & & \\ & \multicolumn{3}{c|}{\rule[0mm]{4cm}{0.1pt}} &\multicolumn{3}{c|}{\rule[0mm]{4cm}{0.1pt}} & & & \\ &\multicolumn{3}{c}{erg s$^{-1}$} &\multicolumn{3}{c}{erg s$^{-1}$} & & & \\ \hline J0204$-$0944 &$-$ &45.38 &$-$ &$-$ &46.45 &$-$ &$-$ &0.09 &$-$\\ J0210$+$0118 &$-$ &45.94 &$-$ &$-$ &47.05 &$-$ &$-$ &0.08 &$-$\\ J0754$+$3033 &$-$ &45.94 &45.66 &$-$ &47.04 &47.29 &$-$ &0.08 &0.02\\ J0754$+$4316 &$-$ &$-$ &45.48 &$-$ &$-$ &47.63 &$-$ &$-$ &0.01\\ J0801$+$4736 &$-$ &$-$ &43.92 &$-$ &$-$ &46.33 &$-$ &$-$&0.004\\ J0809$+$2912 &$-$ &46.25 &$-$ &$-$ &47.16 &$-$ &$-$ &0.12&$-$\\ J0812$+$3031 &$-$ &45.49 &$-$ &$-$ &46.42 &$-$ &$-$ &0.12&$-$\\ J0819$+$0549 &45.29 &44.86 &$-$ &47.49 &46.66 &$-$ &$-$ &0.02&$-$\\ J0842$+$2147 &$-$ &45.51 &$-$ &$-$ &46.49 &$-$ &$-$ &0.11&$-$\\ J0902$+$5707 &45.97 &45.67 &$-$ &47.08 &46.88 &$-$ &$-$ &0.06&$-$\\ J0918$+$2325 &$-$ &45.85 &45.48 &$-$ &47.11 &47.39 &$-$ &0.06&0.01\\ J0925$+$4004 &$-$ &45.87 &45.55 &$-$ &47.21 &47.54 &$-$ &0.05&0.01\\ J0937$+$2937 &$-$ &45.69 &45.38 &$-$ &46.77 &46.85 &$-$ &0.08&0.03\\ J0944$+$2331 &$-$ &45.90 &$-$ &$-$ &46.92 &$-$ &$-$ &0.10&$-$\\ J0959$+$1216 &$-$ &45.58 &$-$ &$-$ &46.82 &$-$ &$-$ &0.06&$-$\\ J1020$+$0447 &$-$ &45.26 &$-$ &$-$ &46.83 &$-$ &$-$ &0.03&$-$\\ J1020$+$3958 &$-$ &45.77 &$-$ &$-$ &47.22 &$-$ &$-$ &0.04&$-$\\ J1030$+$5310 &$-$ &45.90 &$-$ &$-$ &46.76 &$-$ &$-$ &0.14&$-$\\ J1054$+$4152 &$-$ &45.78 &$-$ &$-$ &46.96 &$-$ &$-$ &0.07&$-$\\ J1056$+$4100 &45.34 &45.16 &$-$ &46.83 &46.70 &$-$ &$-$ &0.03&$-$\\ J1145$-$0033 &45.53 &45.51 &$-$ &47.43 &$-$ &$-$ &0.01 &$-$&$-$\\ J1151$+$3355 &$-$ &45.73 &$-$ &$-$ &46.97 &$-$ &$-$ &0.06&$-$\\ J1229$+$3555 &$-$ &45.37 &$-$ &$-$ &46.39 &$-$ &$-$ &0.10&$-$\\ J1304$+$2454 &$-$ &45.92 &$-$ &$-$ &47.01 &47.57 &$-$ &0.08&$-$\\ J1321$+$3741 &$-$ &45.64 &$-$ &$-$ &47.25 &$-$ &$-$ &0.03&$-$\\ J1340$+$4232 &$-$ &45.47 &$-$ &$-$ &46.87 &$-$ &$-$ &0.04&$-$\\ J1353$+$2631 &$-$ &45.63 &45.46 &$-$ &46.75 &47.54 &$-$ &0.08&0.01\\ J1410$+$2955 &$-$ &45.65 &$-$ &$-$ &47.20 &$-$ &$-$ &0.03&$-$\\ J1427$+$2632 &$-$ &$-$ &45.67 &$-$ &$-$ &47.59 &$-$ &$-$&0.01\\ J1432$+$1548 &$-$ &45.65 &$-$ &$-$ &46.96 &$-$ &$-$ &0.05&$-$\\ J1723$+$3417 &$-$ &45.39 &45.73 &$-$ &$-$ &46.66 &$-$ &$-$&0.12\\ J2344$-$0032 &$-$ &45.723 &$-$ &$-$ &46.78 &$-$ &$-$ &0.09&$-$\\ \hline \end{tabular} \end{minipage}\\ \end{table*} \begin{table*} \centering \begin{minipage}{140mm} \caption{Optical luminosity and accretion rate for small-size radio quasars.} \begin{tabular}{@{}lccccccccc@{}} \hline IAU & &log(L$_{bol})$ & & & log(L$_{Edd}$) & &\.m1350 &\.m3000 & \.m5100\\ name &1350\AA &3000\AA & 5100\AA & CIV & MgII & $H\beta$ & & & \\ & \multicolumn{3}{c|}{\rule[0mm]{4cm}{0.1pt}} &\multicolumn{3}{c|}{\rule[0mm]{4cm}{0.1pt}} & & & \\ &\multicolumn{3}{c|}{erg s$^{-1}$} & \multicolumn{3}{c|}{erg s$^{-1}$} & & & \\ \hline J0034$+$0118 &$-$ &45.35 &$-$ &$-$ &46.87 &$-$ &$-$ &0.03 &$-$\\ J0051$-$0902 &$-$ &45.58 &$-$ &$-$ &47.10 &$-$ &$-$ &0.03 &$-$\\ J0130$-$0135 &$-$ &45.83 &$-$ &$-$ &47.19 &$-$ &$-$ &0.05 &$-$\\ J0245$+$0108 &46.05 &45.73 &$-$ &47.23 &47.13 &$-$ &0.07 &0.04 &$-$\\ J0745$+$3142 &$-$ &46.51 &46.26 &$-$ &47.41 &47.78 &$-$ &0.13 &0.03\\ J0811$+$2845 &46.06 &45.85 &$-$ &47.00 &47.08 &$-$ &0.11 &0.06 &$-$\\ J0814$+$3237 &$-$ &45.51 &$-$ &$-$ &46.46 &45.89 &$-$ &0.11 &$-$\\ J0817$+$2237 &$-$ &45.98 &$-$ &$-$ &46.99 &$-$ &$-$ &0.10 &$-$\\ J0828$+$3935 &$-$ &45.38 &$-$ &$-$ &46.62 &$-$ &$-$ &0.06 &$-$\\ J0839$+$1921 &46.33 &46.06 &$-$ &46.95 &46.84 &$-$ &0.24 &0.17 &$-$\\ J0904$+$2819 &$-$ &46.21 &$-$ &$-$ &46.93 &$-$ &$-$ &0.19 &$-$\\ J0906$+$0832 &45.82 &45.56 &$-$ &47.18 &46.84 &$-$ &0.04 &0.05 &$-$\\ J0924$+$3547 &$-$ &45.83 &$-$ &$-$ &46.95 &$-$ &$-$ &0.08 &$-$\\ J0925$+$1444 &$-$ &45.81 &$-$ &$-$ &46.78 &$-$ &$-$ &0.11 &$-$\\ J0935$+$0204 &$-$ &46.03 &45.72 &$-$ &47.28 &47.34 &$-$ &0.06 &0.02\\ J0941$+$3853 &$-$ &45.66 &45.41 &$-$ &47.07 &47.62 &$-$ &0.04 &0.01\\ J0952$+$2352 &$-$ &46.080 &$-$ &$-$ &46.85 &$-$ &$-$ &0.17 &$-$\\ J1000$+$0005 &$-$ &45.42 &$-$ &$-$ &46.38 &$-$ &$-$ &0.11 &$-$\\ J1004$+$2225 &$-$ &45.61 &$-$ &$-$ &46.79 &$-$ &$-$ &0.07 &$-$\\ J1005$+$5019 &46.21 &45.75 &$-$ &46.74 &46.90 &$-$ &0.30 &0.07 &$-$\\ J1006$+$3236 &$-$ &45.34 &$-$ &$-$ &46.44 &$-$ &$-$ &0.08 &$-$\\ J1009$+$0529 &$-$ &46.18 &$-$ &$-$ &47.45 &$-$ &$-$ &0.05 &$-$\\ J1010$+$4132 &$-$ &46.40 &46.00 &$-$ &46.81 &46.81 &$-$ &0.39 &0.15\\ J1023$+$6357 &$-$ &46.20 &$-$ &$-$ &47.16 &$-$ &$-$ &0.11 &$-$\\ J1100$+$1046 &$-$ &45.53 &45.20 &$-$ &46.84 &45.43 &$-$ &0.05 &0.59\\ J1100$+$2314 &$-$ &45.96 &45.87 &$-$ &47.32 &48.05 &$-$ &0.04 &0.01\\ J1107$+$0547 &45.87 &45.54 &$-$ &47.09 &46.67 &$-$ &0.06 &0.07 &$-$\\ J1107$+$1628 &$-$ &46.31 &45.99 &$-$ &46.98 &47.06 &$-$ &0.21 &0.09\\ J1110$+$0321 &$-$ &45.57 &$-$ &$-$ &46.41 &$-$ &$-$ &0.14 &$-$\\ J1118$+$3828 &$-$ &45.64 &45.12 &$-$ &46.69 &48.01 &$-$ &0.09 &0.001\\ J1119$+$3858 &$-$ &45.65 &45.37 &$-$ &47.14 &47.94 &$-$ &0.03 &0.003\\ J1158$+$6254 &$-$ &46.27 &46.07 &$-$ &47.47 &48.12 &$-$ &0.06 &0.01\\ J1217$+$1019 &46.06 &45.47 &$-$ &46.76 &46.60 &$-$ &0.20 &0.08 &$-$\\ J1223$+$3707 &$-$ &45.40 &45.24 &$-$ &46.79 &47.64 &$-$ &0.04 &0.004\\ J1236$+$1034 &$-$ &45.82 &45.56 &$-$ &46.78 &47.82 &$-$ &0.11 &0.01\\ J1256$+$1008 &$-$ &45.44 &$-$ &$-$ &46.46 &45.70 &$-$ &0.10 &$-$\\ J1319$+$5148 &$-$ &46.29 &$-$ &$-$ &47.03 &$-$ &$-$ &0.18 &$-$\\ J1334$+$5501 &$-$ &45.83 &$-$ &$-$ &47.10 &$-$ &$-$ &0.05 &$-$\\ J1358$+$5752 &$-$ &46.39 &$-$ &$-$ &47.20 &$-$ &$-$ &0.16 &$-$\\ J1425$+$2404 &$-$ &46.03 &45.78 &$-$ &47.21 &47.30 &$-$ &0.07 &0.03\\ J1433$+$3209 &$-$ &44.79 &$-$ &$-$ &46.40 &$-$ &$-$ &0.03 &$-$\\ J1513$+$1011 &46.59 &46.20 &$-$ &47.15 &47.05 &$-$ &0.27 &0.14 &$-$\\ J1550$+$3652 &45.95 &45.46 &$-$ &46.76 &46.78 &$-$ &0.15 &0.05 &$-$\\ J1557$+$0253 &45.85 &45.43 &$-$ &46.06 &46.20 &$-$ &0.61 &0.17 &$-$\\ J1557$+$3304 &$-$ &45.74 &$-$ &$-$ &47.01 &$-$ &$-$ &0.05 &$-$\\ J1622$+$3531 &$-$ &45.63 &$-$ &$-$ &46.77 &$-$ &$-$ &0.07 &$-$\\ J1623$+$3419 &45.46 &45.17 &$-$ &46.62 &46.64 &$-$ &0.07 &0.03 &$-$\\ J2335$-$0927 &46.04 &44.77 &$-$ &46.43 &46.18 &$-$ &0.41 &0.04 &$-$\\ \hline \end{tabular} \end{minipage}\\ \end{table*}
\section{Introduction} Given a Poisson variety $\mathfrak{N}$ over $\mathbb{C}$, the {\bf degree zero Poisson homology group} $H\! P_0(\mathfrak{N})$ is defined to be the quotient of $\mathbb{C}[\mathfrak{N}]$ by the linear span of all brackets. If $\mathfrak{N}$ is affine and symplectic, then $H\! P_0(\mathfrak{N})$ is isomorphic to $\H^{\dim\mathfrak{N}}(\mathfrak{N})$ via the map that takes the class of a function to the de Rham class of that function times the appropriate power of the symplectic form. The next interesting case is when $\mathfrak{N}$ is an affine cone that admits a projective symplectic resolution $\mathfrak{M}$. In this case, we may deform the map $\mathfrak{M}\to\mathfrak{N}$ to a map $\mathscr{M}\to\mathscr{N}$, where $\mathscr{M}$ and $\mathscr{N}$ are varieties over the base $\H^2(\mathfrak{M})$ with zero fibers $\mathscr{M}_0=\mathfrak{M}$ and $\mathscr{N}_0=\mathfrak{N}$ \cite{Namiaff, NamiaffII}. Over a generic element $\lambda\in\H^2(\mathfrak{M})$, the map from $\mathscr{M}_\lambda$ to $\mathscr{N}_\lambda$ is an isomorphism of affine varieties. Then $H\! P_0(\mathscr{N})$ is a module over $\mathbb{C}[\H^2(\mathfrak{M})]$ whose specialization at $\lambda$ is isomorphic to $H\! P_0(\mathscr{N}_\lambda)$. When $\lambda$ is generic, this is isomorphic to $\H^{\dim\mathscr{N}_\lambda}(\mathscr{N}_\lambda)\cong \H^{\dim\mathscr{M}_\lambda}(\mathscr{M}_\lambda)\cong\H^{\dim\mathfrak{M}}(\mathfrak{M})$. Etingof and Schedler conjecture that the dimension of the zero fiber $H\! P_0(\mathfrak{N})$ is also equal to that of $\H^{\dim\mathfrak{M}}(\mathfrak{M})$; equivalently, they conjecture that $H\! P_0(\mathscr{N})$ is free over $\mathbb{C}[\H^2(\mathfrak{M})]$ \cite[1.3.1(a)]{ES11}. They prove this conjecture for the Springer resolution and for Hilbert schemes of points on ALE spaces. The goal of this paper is to both strengthen and prove this conjecture for hypertoric varieties, and to pose an analogous strengthening for other projective symplectic resolutions of affine cones. A hypertoric variety is an affine cone $\mathfrak{N}$ that admits a projective symplectic resolution $\mathfrak{M}$, equivariant for an effective Hamiltonian action of a torus $T$, with $\dim T = \frac 1 2 \dim\mathfrak{N}$. A hypertoric variety $\mathfrak{N}$ comes with a ``dual" hypertoric variety $\mathfrak{N}^!$, equipped with an action of its own torus $T^!$; the relationship between these dual pairs has been studied in \cite{GDKD} and \cite{BLPWtorico}. One of the first properties of a dual pair is that the cohomology group $\H^2(\mathfrak{M})$ is canonically isomorphic to the Lie algebra of $T^!$. Our main result (Theorem \ref{main}) states that $H\! P_0(\mathscr{N})$ is isomorphic as a graded module over $\mathbb{C}[\H^2(\mathfrak{M})]\cong \H^*(BT^!)$ to the equivariant intersection cohomology group $I\! H^*_{T^!}(\mathfrak{N}^!)$, where the grading in Poisson homology is induced by the conical action of the multiplicative group. In particular, this implies that $H\! P_0(\mathscr{N})$ is a free module over $\mathbb{C}[\H^2(\mathfrak{M})]$. The relationship between a dual pair of hypertoric varieties is a special case of a relationship between pairs of projective symplectic resolutions called {\bf symplectic duality}, which is being studied by Braden, Licata, Webster, and the author (see \cite[1.5]{BLPWtorico} and \cite{BLPWgco}). Examples of symplectic dual pairs, along with a conjectural extension of Theorem \ref{main} to this setting (Conjecture \ref{general dual}), are given in Remark \ref{sd}. We also note that Poisson homology of Poisson varieties is closely related to Hochschild homology of their quantizations. More precisely, let $A$ be a quantization of $\mathbb{C}[\mathscr{N}]$, and let $H\! H_0(A)$ be the quotient of $A$ by the linear span of all commutators. Then $H\! H_0(A)$ is a filtered vector space such that $\grH\! H_0(A)$ admits a canonical map from $H\! P_0(\mathscr{N})$; Etingof and Schedler conjecture that this map is an isomorphism \cite[1.3.3]{ES11}. In Remark \ref{quantization} and Conjecture \ref{nc-dual}, we conjecture the appropriate analogue of Theorem \ref{main} and Conjecture \ref{general dual} in the quantized setting. The paper is organized as follows: In Section \ref{hypertoric} we give a basic construction of a hypertoric variety and its dual. Section \ref{results} is devoted to the statement of the main theorem and associated conjectures. We give a combinatorial presentation of $H\! P_0(\mathscr{N})$ in Section \ref{presentation}, which we use in Section \ref{numerics} to prove that the main theorem holds on the numerical level (that is, we prove that the modules are isomorphic without obtaining a canonical isomorphism). Section \ref{canonical}, which draws heavily on the machinery of \cite{TP08}, establishes the canonical isomorphism.\\ \noindent {\em Acknowledgments:} The authors is grateful to Travis Schedler for introducing him to Poisson homology and explaining the various conjectures in \cite{ES11}. \section{Hypertoric varieties}\label{hypertoric} Fix a positive integer $n$, and let $V = \mathbb{C}^n$. Let $T^n := (\C^*)^n$ be the coordinate torus acting on $V$ in the standard way, and let $X(T^n) \cong \mathbb{Z}^n$ be its character lattice. The vector space $V\oplus V^*$ carries a natural symplectic form, and the induced action of $T^n$ is Hamiltonian with moment map $$\Phi:V\oplus V^*\to X(T^n)_\mathbb{C}\cong\mathbb{C}^n$$ given by the formula $\Phi(z,w) = (z_1w_1,\ldots,z_nw_n)$. Let $\iota:G\hookrightarrow T^n$ be a connected algebraic subtorus, and let $\iota^*:X(T^n)\to X(G)$ be the pullback map on characters. Assume additionally that $\iota$ is {\bf unimodular}, which means that for some (equivalently any) choice of basis for $X(G)$, all minors of $\iota^*$ belong to the set $\{-1, 0, 1\}$. The action of $G$ on $V\oplus V^*$ is Hamiltonian with moment map $$\mu = \iota^*_\mathbb{C}\circ\Phi:V\oplus V^*\to X(G)_\mathbb{C}.$$ Fix once and for all a generic character $\theta\in X(G)$. Let $$\mathscr{M} := V\oplus V^*/\!\!/\!_\theta\, G = \operatorname{Proj}\mathbb{C}[z_1,\ldots,z_n,w_1,\ldots,w_n,t]^G$$ and $$\mathscr{N} := V\oplus V^*/\!\!/\!_0\, G = \operatorname{Spec}\mathbb{C}[z_1,\ldots,z_n,w_1,\ldots,w_n]^G,$$ where $G$ acts on $t$ with weight $\theta$. The map $\mu$ descends to a pair of maps $$\pi:\mathscr{M}\to X(G)_\mathbb{C}\and \bar\pi:\mathscr{N}\to X(G)_\mathbb{C},$$ and we let $$\mathfrak{M} := \pi^{-1}(0)\and \mathfrak{N} := \bar\pi^{-1}(0).$$ These two spaces are called {\bf hypertoric varieties} \cite{BD,Pr07}. The variety $\mathfrak{N}$ is an affine cone, and $\mathfrak{M}$ is a projective symplectic resolution of $\mathfrak{N}$. Let $T := T^n/G$; the action of $T^n$ on $V\oplus V^*$ descends to an effective Poisson action of $T$ on all of the aforementioned spaces. There is an additional (non-Poisson) action of $S\cong \C^*$ induced by the inverse scalar action on $V\oplus V^*$. We have the following $S\times T$-equivariant commutative diagram, where $T$ fixes $X(G)_\mathbb{C}$ and $S$ acts on $X(G)_\mathbb{C}$ with weight $-2$. \[\xymatrix{\mathfrak{M} \ar[d]\ar[rr] & & \mathscr{M} \ar[d]\ar[rr]^{\pi} & & X(G)_\mathbb{C}\ar[d]^{=} \\ \mathfrak{N} \ar[rr] & & \mathscr{N} \ar[rr]^{\bar\pi} & & X(G)_\mathbb{C} }\] The vector space $X(G)_\mathbb{C}$ surjects onto $\H^2(\mathfrak{M})$ via the Kirwan map, and $\mathscr{M}$ is the pullback of the universal Poisson deformation of $\mathfrak{M}$ \cite{Namiaff, NamiaffII}. \begin{remark} Some justification is needed for the fact that in the introduction $\mathscr{M}$ and $\mathscr{N}$ were families over $\H^2(\mathfrak{M})$, whereas now we define them as families over $X(G)_\mathbb{C}$. In most cases, the Kirwan map from $X(G)_\mathbb{C}$ to $\H^2(\mathfrak{M})$ is an isomorphism, so the two definitions agree; this fails if and only if the image of the Lie algebra map $\mathfrak{g}\to\mathfrak{t}^n\cong\mathbb{C}^n$ contains a coordinate line. (The kernel of the Kirwan map in all degrees is computed in different ways in \cite[2.4]{Ko}, \cite[1.1]{HS}, and \cite[3.2.2]{Pr07}.) In those cases where we do not get an isomorphism, we really need the family over $X(G)_\mathbb{C}$ rather than over $\H^2(\mathfrak{M})$ for our results to hold as stated. \end{remark} \begin{remark}\label{almost universal} Even if the Kirwan map is an isomorphism in degree 2, it may or may not be the case that $\mathscr{N}$ is the universal Poisson deformation of $\mathfrak{N}$. In general, this universal deformation is parameterized by $\H^2(\mathfrak{M})/W$, where $W$ is the Namikawa Weyl group of $\mathfrak{N}$ (which may or may not be trivial). The deformation $\mathscr{N}$ is the pullback of the universal deformation from $\H^2(\mathfrak{M})/W$ to $X(G)_\mathbb{C}$. We will refer to this as the {\bf quasi-universal} deformation of $\mathfrak{N}$. \end{remark} Let $(T^n)^*$ be the dual torus to $T^n$, and let $G^!\subset (T^n)^*$ be the inclusion of the connected subtorus whose Lie algebra is perpendicular to that of $G\subset T^n$. Fixing a generic character $\theta^!\in X(G^!)$, we may construct new spaces \[\xymatrix{\mathfrak{M}^! \ar[d]\ar[rr] & & \mathscr{M}^! \ar[d]\ar[rr]^{\pi^!} & & X(G^!)_\mathbb{C}\ar[d]^{=} \\ \mathfrak{N}^! \ar[rr] & & \mathscr{N}^! \ar[rr]^{\bar\pi^!} & & X(G^!)_\mathbb{C} }\] as above. This diagram will be $S\times T^!$-equivariant, where $S$ is as before and $T^! := (T^n)^*/G^!$. The relationship between $\mathfrak{N}$ and $\mathfrak{N}^!$ (and all of their associated geometry and representation theory) was explored in detail in \cite{GDKD} and \cite{BLPWtorico} (see also Remark \ref{sd}). \section{Results and conjectures}\label{results} For any $\lambda\in X(G)_\mathbb{C}\cong\mathfrak{g}^*$, let $$\mathscr{M}_\lambda := \pi^{-1}(\lambda)\and \mathscr{N}_\lambda := \bar\pi^{-1}(\lambda).$$ The degree zero Poisson homology group $H\! P_0(\mathscr{N})$ is a module over $\mathbb{C}[X(G)_\mathbb{C}] \cong \operatorname{Sym}\mathfrak{g}$, and for any $\lambda$, we have $$H\! P_0(\mathscr{N}_\lambda) \cong H\! P_0(\mathscr{N})\otimes_{\operatorname{Sym}\mathfrak{g}}\mathbb{C}_{\lambda}.$$ The action of $S$ induces positive integer gradings of $H\! P_0(\mathscr{M})$ and $\operatorname{Sym}\mathfrak{g}$, with $\mathfrak{g}$ sitting in degree 2. This grading descends to $H\! P_0(\mathfrak{N}) = H\! P_0(\mathscr{N}_0)$, but not to $H\! P_0(\mathscr{N}_\lambda)$ for nonzero $\lambda$. Observe that $T^!$ is canonically dual to $G$; in particular, $\H^*(BT^!)$ is canonically isomorphic as a graded ring to $\operatorname{Sym}\mathfrak{g}$. Our main result is the following. \begin{theorem}\label{main} There is a canonical isomorphism of graded $\operatorname{Sym}\mathfrak{g}$-modules $H\! P_0(\mathscr{N}) \cong I\! H^*_{T^!}(\mathfrak{N}^!)$. In particular, $H\! P_0(\mathfrak{N})$ is isomorphic as a graded vector space to $I\! H^*(\mathfrak{N}^!)$. \end{theorem} \begin{remark}\label{freeness} A combinatorial interpretation of the intersection cohomology Betti numbers of $\mathfrak{N}^!$ was given in \cite[4.3]{PW07}. In particular, they vanish in odd degree, thus $I\! H^*_{T^!}(\mathfrak{N}^!)$ is a free $\operatorname{Sym} \mathfrak{g}$-module. Theorem \ref{main} is therefore a strengthening in the hypertoric case a conjecture of Etingof and Schedler \cite[1.3.1(a)]{ES11}, which says (for an affine cone that admits a projective symplectic resolution) that the degree zero Poisson homology of the quasi-universal deformation is a free module over the coordinate ring of the base. In fact, proving freeness (Corollary \ref{free}) is one of the steps toward proving Theorem \ref{main}. \end{remark} \begin{remark}\label{sd} The relationship between $\mathfrak{N}$ and $\mathfrak{N}^!$ is a special case of a phenomenon called {\bf symplectic duality} \cite[1.5]{BLPWtorico}, which relates pairs of affine cones that admit projective symplectic resolutions. Theorem \ref{main} invites the following generalization. \begin{conjecture}\label{general dual} Let $\mathfrak{N}$ and $\mathfrak{N}^!$ be a symplectic dual pair of cones $\mathfrak{N}$ and $\mathfrak{N}^!$ admitting projective symplectic resolutions $\mathfrak{M}$ and $\mathfrak{M}^!$. Let $\mathscr{N}$ be the quasi-universal deformation of $\mathfrak{N}$, and let $T^!$ be a maximal torus in the Hamiltonian automorphism group of $\mathfrak{M}^!$. There exist canonical isomorphisms of graded vector spaces $$H\! P_0(\mathfrak{N})\congI\! H^*(\mathfrak{N}^!) \and H\! P_0(\mathscr{N})\congI\! H^*_{T^!}(\mathfrak{N}^!),$$ where the second isomorphism is compatible with the module structure over $\mathbb{C}[\H^2(\mathfrak{M})]\cong \H^*(BT^!)$. \end{conjecture} Note that $I\! H^*(\mathfrak{N}^!)\subset \H^*(\mathfrak{M}^!)$ is always concentrated in even degree \cite[2.5]{BLPWquant}, and therefore the deformation coming from equivariant cohomology is always free. Thus Conjecture \ref{general dual} would imply the conjecture of Etingof and Schedler from Remark \ref{freeness} for any cone that has a symplectic dual. While a general definition of a symplectic dual pair is still in preparation \cite{BLPWgco}, we already know a number of pairs of cones that should be examples (in addition to the hypertoric examples in this paper). For instance: \begin{itemize} \item If $G$ is a simple algebraic group and $G^L$ is its Langlands dual, then the nilpotent cone of $\mathfrak{g}$ should be dual to the nilpotent cone of $\mathfrak{g}^L$. In this case, both Poisson homology and intersection cohomology are 1-dimensional \cite[1.6]{ES10}. \item A normal slice inside the nilpotent cone to a subregular niloptent orbit in a simply-laced simple Lie algebra $\mathfrak{g}$ should be dual to the closure of the minimal nontrivial nilpotent orbit in $\mathfrak{g}^L\cong\mathfrak{g}$. The Poisson homology Poincar\'e polynomial of the slice is computed in \cite{AL98} and the intersection cohomology Poincar\'e polynomial of the orbit closure is computed in \cite[6.4.2]{MOV03}, and they agree.\footnote{In the arXiv version \cite{MOV03}, it is 6.2.2.} \item If $G$ and $G'$ are any two simply-laced simple algebraic groups and $\Gamma$ and $\Gamma'$ are the corresponding finite subgroups of $\operatorname{SL}_2\mathbb{C}$, then the affinization of the moduli space of $G$-instantons on a crepant resolution of $\mathbb{C}^2/\Gamma'$ should be dual to the affinization of the moduli space of $G'$-instantons on a crepant resolution of $\mathbb{C}^2/\Gamma$. The previous example is a special case where $G'$ and $\Gamma'$ are trivial. \end{itemize} \end{remark} \begin{remark}\label{quantization} Let $A$ be a quantization of $\mathscr{N}$; that is, a filtered algebra whose associated graded ring is isomorphic to $\mathbb{C}[\mathscr{N}]$ (with grading induced by the $S$-action), inducing the given Poisson structure. Then $A$ has a central quotient $A_0$ which is a quantization of $\mathfrak{N}$. The Hochschild homology group $H\! H_0(A) := A\big{/}[A,A]$ admits a filtration whose associated graded module admits a surjection from $H\! P_0(\mathscr{N})$ as a graded module over $Z(A) \cong \operatorname{Sym}\mathfrak{g}$; in particular, we also get a surjection from $H\! P_0(\mathfrak{N})$ to $H\! H_0(A_0)$. If the conjecture \cite[1.3.1(a)]{ES11} holds, then these surjections are both isomorphisms \cite[\S 1.3]{ES11}. The appropriate analogue of Conjecture \ref{general dual} is the following. \begin{conjecture}\label{nc-dual} In the situation of Conjecture \ref{general dual}, there exist canonical isomorphisms of filtered vector spaces $$H\! H_0(A_0)\congI\! H^*_S(\mathfrak{N}^!)\otimes_{\mathbb{C}[u]}\mathbb{C}_1 \and H\! H_0(A)\congI\! H^*_{S\times T^!}(\mathfrak{N}^!)\otimes_{\mathbb{C}[u]}\mathbb{C}_1,$$ where $\mathbb{C}_1$ is the one-dimensional module over $\H^*(BS)\cong\mathbb{C}[u]$ annihilated by $u-1$. The second isomorphism is compatible with the module structure over $Z(A)\cong \H^*(BT^!)$. \end{conjecture} In the hypertoric case, we make take $A$ to be the ring of $G$-invariant differential operators on $V$, which is called the {\bf hypertoric enveloping algebra} \cite[5.2]{BLPWtorico}. Since we know by Theorem \ref{main} that $H\! P_0(\mathscr{N})$ is a free module, $H\! H_0(A)$ must be a filtered free module whose associated graded module is canonically isomorphic to $H\! P_0(\mathscr{N})$. It should be possible to establish the canonical isomorphism of Conjecture \ref{nc-dual} using techniques similar to those that we use to prove Theorem \ref{main}; the main task would be to extend the techniques of \cite{TP08} to a setting that takes the $S$-action into account. \end{remark} \begin{remark}\label{duals} Let $\lambda\in X(G)_\mathbb{C}$ be generic. In this case $\mathscr{N}_\lambda$ is smooth and affine, thus \begin{equation*}\label{deformation} H\! P_0(\mathscr{N})\otimes_{\operatorname{Sym}\mathfrak{g}}\mathbb{C}_\lambda \congH\! P_0(\mathscr{N}_\lambda) \cong \H^{\dim \mathscr{N}_\lambda}(\mathscr{N}_\lambda) \cong \H^{\dim \mathscr{M}_\lambda}(\mathscr{M}_\lambda) \cong \H^{\dim \mathfrak{M}}(\mathfrak{M}). \end{equation*} Here the second isomorphism is given by multiplication by the appropriate power of the symplectic form, the third isomorphism comes from the fact that $\mathscr{N}_\lambda$ is isomorphic to $\mathscr{M}_\lambda$, and the fourth isomorphism comes from the topological triviality of the family $\pi:\mathscr{M}\to X(G)_\mathbb{C}$. On the other hand, \cite[7.21]{BLPWtorico} implies that $I\! H^*_{T^!}(\mathfrak{N}^!)\otimes_{\operatorname{Sym}\mathfrak{g}}\mathbb{C}_\lambda$ is canonically {\em dual} to $\H^{\dim \mathfrak{M}}(\mathfrak{M})$. Comparing these two results means gives us a nondegenerate inner product on the vector space $\H^{\dim \mathfrak{M}}(\mathfrak{M})$. Somewhat surprisingly, this pairing depends nontrivially on $\lambda$. Indeed, given a generic class $\lambda\in X(G)_\mathbb{C}$ with image $\bar\lambda\in\H^2(\mathfrak{M})$, it is straightforward to check that the pairing of the class $\bar\lambda^{\frac{1}{2}\dim\mathfrak{M}}\in\H^{\dim \mathfrak{M}}(\mathfrak{M})$ with itself is equal to 1. Moreover, for any nonzero complex number $c$, the pairing associated to $c\lambda$ is equal to $c^{-\dim\mathfrak{M}}$ times the pairing associated to $\lambda$. It would be interesting to understand this family of inner products in more detail. An analogue of \cite[7.21]{BLPWtorico} is part of the package for all symplectic dual pairs, so a similar phenomenon should arise for all of the examples mentioned in Remark \ref{sd}. \end{remark} \section{A presentation of \boldmath{$H\! P_0(\mathscr{N})$}}\label{presentation} For each $\a\in X(T^n) \subset (\mathfrak{t}^n)^*$, consider the differential operator $$\partial_\a: \operatorname{Sym}\mathfrak{t}^n\to\operatorname{Sym}\mathfrak{t}^n$$ defined by putting $\partial_\a x = \langle \a,x\rangle$ for all $x\in\mathfrak{t}^n$ and extending via the Leibniz rule. It will be convenient for us to work in coordinates; we identify $X(T^n)$ with $\mathbb{Z}^n$ and $\operatorname{Sym}\mathfrak{t}^n$ with $\mathbb{C}[e_1,\ldots,e_n]$, and we have $\partial_\a e_i = \a_i$. We will be particularly interested in those operators $\partial_\a$ for which $\a$ is in the kernel of $\iota^*:X(T^n)\to X(G)$. Consider the vector space $$J:= \mathbb{C}\big\{\partial_\a e^\b\mid \a\in\ker\iota^*,\; \b\in\mathbb{N}^n,\; \operatorname{Supp}(\a)\subset\operatorname{Supp}(\b)\big\}\subset\mathbb{C}[e_1,\ldots,e_n],$$ where the {\bf support} of an element of $\mathbb{Z}^n$ is the set of coordinates where that element is nonzero. This is not an ideal, but it is a module over $$\operatorname{Sym}\mathfrak{g}\subset\operatorname{Sym}\mathfrak{t}^n = \mathbb{C}[e_1,\ldots,e_n]$$ because $\partial_\a x = 0$ for all $x\in\mathfrak{g}\subset\mathfrak{t}^n$. \begin{remark} For any $\a$ and $\b$, we have $\partial_{\a}+\partial_\b = \partial_{\a+\b}$. For this reason, we may restrict our attention in the definition of $J$ to those $\a$ which are primitive and have minimal support. The unimodularity condition implies that for such an $\a$, $\a_i\in\{-1,0,1\}$ for all $i$. Such an $\a$ is called a {\bf signed circuit}, and there are only finitely many of them. \end{remark} \begin{proposition}\label{explicit} $H\! P_0(\mathscr{N})$ is isomorphic to $\mathbb{C}[e_1,\ldots,e_n]/J$ as a graded $\operatorname{Sym}\mathfrak{g}$-module. \end{proposition} \begin{proof} Recall that $H\! P_0(\mathscr{N})$ is defined as the quotient of $\mathbb{C}[\mathscr{N}] = \mathbb{C}[z_1,\ldots,z_n,w_1,\ldots,w_n]^G$ by the linear span of all brackets, and consider the graded $\operatorname{Sym}\mathfrak{g}$-module homomorphism $$\psi:\mathbb{C}[e_1,\ldots,e_n]\toH\! P_0(\mathscr{N})$$ taking $e_i$ to the class represented by $z_iw_i$. We will show that $\psi$ is surjective with kernel $J$. The invariant ring $\mathbb{C}[z_1,\ldots,z_n,w_1,\ldots,w_n]^G$ consists of all monomials $z^\b w^\d$ with $\b,\d\in\mathbb{N}^n$ and $\iota^*(\b-\d) = 0$. The Poisson bracket is given by the formula $$\{p(z,w), q(z,w)\} = \sum_{i=1}^n r_i(z,w),$$ where $r_i(z,w)$ is the coefficient of $dz_i\wedge dw_i$ in the expansion of $dp(z,w)\wedge dq(z,w)$. In particular, we have $$\{z_iw_i, z^{\b}w^{\d}\}= (\b_i-\d_i)z^{\b}w^{\d}.$$ This tells us that the class of $z^{\b}w^{\d}$ in $H\! P_0(\mathscr{N})$ is zero unless $\b=\d$, and therefore that $\psi$ is surjective. The remaining relations in $H\! P_0(\mathscr{N})$ come from brackets of the form $$\left\{(zw)^\gamma z^\b w^\d, (zw)^\epsilon z^\d w^\b\right\}$$ for some $\b,\d,\gamma,\epsilon\in\mathbb{N}^n$ with $\iota^*(\b-\d) = 0$. This bracket expands to $$\sum_{i=1}^n (\b_i-\gamma_i)(\b_i+\gamma_i+\d_i+\epsilon_i) (zw)^{\b+\d+g+\epsilon}/(z_iw_i) = \psi\left(\partial_{\b-\gamma}\, e^{\b+\gamma+\d+\epsilon}\right).$$ Thus the kernel of $\psi$ is contained in $J$. The fact that $J$ is contained in the kernel follows from unimodularity: given a signed circuit $\a$, we can find a unique pair $\b,\gamma\in\mathbb{N}^n$ such that $\a=\b-\gamma$ and $\operatorname{Supp}(\a) = \operatorname{Supp}(\b)\sqcup\operatorname{Supp}(\gamma)$. Then every monomial whose support contains that of $\a$ is a multiple of $e^{\b+\gamma}$. \end{proof} \begin{remark} If the condition of unimodularity is dropped, then the last sentence of the proof of Proposition \ref{explicit} will fail, and this will cause Theorem \ref{main} to fail, as well. Geometrically, unimodularity ensures that $\mathfrak{M}$ is a manifold rather than an orbifold. Thus Conjecture \ref{sd} is really about affine cones that admit symplectic resolutions; orbifold resolutions are not good enough. \end{remark} \section{Numerics}\label{numerics} In this section we will prove Theorem \ref{main} minus the word ``canonical". That is, we will show that $H\! P_0(\mathscr{N})$ is a free $\operatorname{Sym}\mathfrak{g}$-module with the same Hilbert series as $I\! H^*_{T^!}(\mathfrak{N}^!)$. This will be a necessary first step toward establishing the canonical isomorphism, which we will do in the next section. \begin{lemma}\label{degeneration} The graded $\operatorname{Sym}\mathfrak{g}$-module $\mathbb{C}[e_1,\ldots,e_n]/J$ degenerates flatly to $\mathbb{C}[e_1,\ldots,e_n]/J_{\Delta^{\operatorname{bc}}}$, where $J_{\Delta^{\operatorname{bc}}}$ is the Stanley-Reisner ideal of the broken circuit complex of the matroid associated to the inclusion $G\subset T^n$. \end{lemma} \begin{proof} Consider the graded lexicographical term order on $\mathbb{C}[e_1,\ldots,e_n]$, which allows us to define the initial $\operatorname{Sym}\mathfrak{g}$-module $$\operatorname{in}(J) := \{\operatorname{in}(f)\mid f\in J\}\subset\mathbb{C}[e_1,\ldots,e_n].$$ We want to show that $\operatorname{in}(J)=J_{\Delta^{\operatorname{bc}}}$. For all $\a\in\ker\iota^*$ and $\b\in\mathbb{N}^n$ with $\operatorname{Supp}(\a)\subset\operatorname{Supp}(\b)$, $\operatorname{in}(\partial_\a e^\b) = \a_i\b_i e^\b/e_i$, where $i$ is the maximal element of the support of $\a$. Thus we have $$\operatorname{in}(J) \supset \mathbb{C}\big\{\operatorname{in}(\partial_\a e^\b)\mid \a\in\ker\iota^*,\; \b\in\mathbb{N}^n,\; \operatorname{Supp}(\a)\subset\operatorname{Supp}(\b)\big\} = J_{\Delta^{\operatorname{bc}}}.$$ We do not yet know whether or not this containment is an equality, because the set of initial terms of a basis for a module need not form a basis for the initial module. What we do know is that $\mathbb{C}[e_1,\ldots,e_n]/\operatorname{in}(J)$ is isomorphic to a quotient of $\mathbb{C}[e_1,\ldots,e_n]/J_{\Delta^{\operatorname{bc}}}$ by some $\operatorname{Sym}\mathfrak{g}$-submodule, which we will call $Q$. By \cite[4.3]{PW07} and \cite[Proposition 1]{PS}, $\mathbb{C}[e_1,\ldots,e_n]/J_{\Delta^{\operatorname{bc}}}$ is a free $\operatorname{Sym}\mathfrak{g}$-module with the same Hilbert series as $I\! H^*_{T^!}(\mathfrak{N}^!)$. As noted in Remark \ref{duals}, for generic $\lambda\in X(G)_\mathbb{C} \cong \mathfrak{g}^*$, $H\! P_0(\mathscr{N})\otimes_{\operatorname{Sym}\mathfrak{g}}\mathbb{C}_\lambda$ is isomorphic to $\H^{\dim\mathfrak{M}}(\mathfrak{M})$, which is in turn dual to $I\! H^*_{T^!}(\mathfrak{N}^!)\otimes_{\operatorname{Sym}\mathfrak{g}}\mathbb{C}_\lambda$. In particular $H\! P_0(\mathscr{N})\otimes_{\operatorname{Sym}\mathfrak{g}}\mathbb{C}_\lambda$ and $I\! H^*_{T^!}(\mathfrak{N}^!)\otimes_{\operatorname{Sym}\mathfrak{g}}\mathbb{C}_\lambda$ have the same vector space dimension, and therefore so do $$\big(\mathbb{C}[e_1,\ldots,e_n]/\operatorname{in}(J)\big)\otimes_{\operatorname{Sym}\mathfrak{g}}\mathbb{C}_\lambda \and \big(\mathbb{C}[e_1,\ldots,e_n]/J_{\Delta^{\operatorname{bc}}}\big)\otimes_{\operatorname{Sym}\mathfrak{g}}\mathbb{C}_\lambda.$$ This implies that $Q\otimes_{\operatorname{Sym}\mathfrak{g}}\mathbb{C}_\lambda = 0$. But since $Q$ is a submodule of a free module and $\lambda$ is generic, this implies that $Q=0$, and we are done. \end{proof} \begin{remark} In the last paragraph of the proof of Lemma \ref{degeneration}, we invoked the statement that $\H^{\dim\mathfrak{M}}(\mathfrak{M})$ is (naturally) dual to $I\! H^*_{T^!}(\mathfrak{N}^!)\otimes_{\operatorname{Sym}\mathfrak{g}}\mathbb{C}_\lambda$. This result, which is established in \cite[7.21]{BLPWtorico}, builds on an enormous amount of background material, and might be frustrating to a reader who does not want to take the time to learn all about hypertoric category $\mathcal{O}$ and symplectic duality. In fact, this was overkill; all we needed to know was that the dimension of $\H^{\dim\mathfrak{M}}(\mathfrak{M})$ is equal to that of $I\! H^*_{T^!}(\mathfrak{N}^!)\otimes_{\operatorname{Sym}\mathfrak{g}}\mathbb{C}_\lambda$, which is the same as the total dimension of $I\! H^*(\mathfrak{N}^!)$. This fact follows from \cite[3.5 \& 4.3]{PW07}, along with the basic combinatorial fact that the sum of the $h$-numbers of $\Delta^{\operatorname{bc}}$ is equal to the top $h$-number of the dual matroid. \end{remark} \begin{corollary}\label{free} $H\! P_0(\mathscr{N})$ is a free $\operatorname{Sym}\mathfrak{g}$-module with the same Hilbert series as $I\! H^*_{T^!}(\mathfrak{N}^!)$. \end{corollary} \begin{proof} This follows from Proposition \ref{explicit} and Lemma \ref{degeneration} along with \cite[4.3]{PW07} and \cite[Proposition 1]{PS}. \end{proof} \section{The canonical isomorphism}\label{canonical} In this section we prove Theorem \ref{main}. Consider the hyperplane arrangement $\cH = \{H_1,\ldots,H_n\}$ in the vector space $\mathfrak{g}\subset\mathfrak{t}^n$ where $H_i$ is the intersection of $\mathfrak{g}$ with the $i^\text{th}$ coordinate hyperplane of $\mathfrak{t}^n\cong\mathbb{C}^n$. (This is the hyperplane arrangement that is standardly associated with the hypertoric variety $\mathfrak{N}^!$, for example in \cite{BD} or \cite{Pr07}.) A {\bf flat} of $\cH$ is a subspace of $\mathfrak{g}$ obtained by intersecting some (possibly empty) subset of the hyperplanes. Let $L_\cH$ be the poset of flats of $\cH$, ordered by reverse inclusion. The set $L_\cH$ has a topology in which $U\subset L_\cH$ is open if and only if whenever $F\leq F'$ and $F'\in U$, we have $F\in U$. If $\mathcal{S}$ is a sheaf on $L_\cH$ and $F$ is a flat, let $\mathcal{S}(F)$ be the stalk of $\mathcal{S}$ at $F$. For each $F$, there is a minimal open set $U_F$ containing $F$, so $\mathcal{S}(F)$ is simply equal to $\mathcal{S}(U_F)$. We have $U_F\subset U_{F'}$ if and only if $F\leq F'$, therefore we have a collection of restriction maps $r(F,F'):\mathcal{S}(F')\to\mathcal{S}(F)$ for every pair of comparable flats. By \cite[1.1]{TP08}, a sheaf is completely determined by its stalks and these restriction maps. Let $\mathcal{A}$ be the sheaf of algebras with $\mathcal{A}(F) = \operatorname{Sym}(\mathfrak{g}/F)$, along with the obvious restriction maps. This sheaf is called the {\bf structure sheaf} of $L_\cH$. We will be interested in sheaves of graded $\mathcal{A}$-modules on $L_\cH$. A sheaf $\cL$ of graded $\mathcal{A}$-modules is called a {\bf minimal extension sheaf} if it satisfies four properties: \begin{itemize} \item $\cL$ is indecomposable \item $\cL$ is flabby \item $\cL(F)$ is a free $\cL(F)$-module for all $F$ \item $\cL(\mathfrak{g}) \cong \mathcal{A}(\mathfrak{g}) = \mathbb{C}$. \end{itemize} Such a sheaf exists by \cite[1.10]{TP08}. For any two minimal extension sheaves $\cL$ and $\cL'$, there exists an isomorphism of $\mathcal{A}$-modules from $\cL$ to $\cL'$, and this isomorphism is unique up to scalar multiplication \cite[2.7]{TP08}. Furthermore, there is a particular minimal extension sheaf $\cL$ (defined by applying a certain localization functor to the equivariant $\operatorname{IC}$-sheaf of $\mathfrak{N}^!$) with the property that $\cL(0) = I\! H^*_{T^!}(\mathfrak{N}^!)$ \cite[2.7]{TP08}. To prove Theorem \ref{main}, we will find another minimal extension sheaf $\mathcal{M}$ with the property that $\mathcal{M}(0)$ is canonically isomorphic to $H\! P_0(\mathscr{N})$. This will get us an isomorphism between $\cL(0) = I\! H^*_{T^!}(\mathfrak{N}^!)$ and $\mathcal{M}(0) \cong H\! P_0(\mathscr{N})$ of modules over $\mathcal{A}(0) = \operatorname{Sym}\mathfrak{g}$. Since the isomorphism between $\cL$ and $\mathcal{M}$ is unique up to scalar multiplication, the isomorphism between stalks at 0 will be canonical up to scalar multiplication. It can then be made completely canonical by noting that both $I\! H^*_{T^!}(\mathfrak{N}^!)$ and $H\! P_0(\mathscr{N})$ are canonically isomorphic to $\mathbb{C}$ in degree zero. Recall that the data with which we began in Section \ref{hypertoric} was an inclusion of tori $G\hookrightarrow T^n$, or equivalently an inclusion of abelian Lie algebras $\mathfrak{g}\to\mathfrak{t}^n$. For each flat $F$, let $$\mathfrak{t}^n_F := \mathfrak{t}^n/\mathbb{C}\{e_i\mid F\not\subset H_i\},$$ and consider the inclusion $\mathfrak{g}/F\hookrightarrow\mathfrak{t}^n_F.$ With this starting point, we can repeat all of the constructions in this paper with $F$ in the subscript. Define two sheaves of graded $\mathcal{A}$-modules $\mathcal{M}$ and $\mathcal{R}^{\operatorname{bc}}$, where $$\mathcal{M}(F):= \operatorname{Sym}\mathfrak{t}^n_F/J_F \and \mathcal{R}^{\operatorname{bc}}(F) := \operatorname{Sym}\mathfrak{t}^n_F/J_{\Delta_F^{\operatorname{bc}}},$$ with the obvious restriction maps. Lemma \ref{degeneration} tells us that $\mathcal{M}$ admits a filtration whose associated graded module is isomorphic to $\mathcal{R}^{\operatorname{bc}}$. We know that $\mathcal{R}^{\operatorname{bc}}$ is a minimal extension sheaf by \cite[3.9]{TP08}, thus the same is true of $\mathcal{M}$. This completes the proof of Theorem \ref{main}. \begin{remark} {\bf What just happened?} We now summarize our approach to the proof of Theorem \ref{main}, in case the central idea got lost in the machinery of \cite{TP08}. Given two free graded $\operatorname{Sym}\mathfrak{g}$-modules $L$ and $M$ with the same Hilbert series, they are necessarily isomorphic, but not canonically so. The problem is that graded $\operatorname{Sym}\mathfrak{g}$-modules are very floppy things--that is, they have lots of automorphisms. On the other hand, a minimal extension sheaf on $L_\cH$ is a rigid object--that is, it has only scalar automorphisms--whose space of global sections is a graded $\operatorname{Sym}\mathfrak{g}$-module. Thus, if $L$ and $M$ can be promoted to minimal extension sheaves $\cL$ and $\mathcal{M}$, then $\cL$ and $\mathcal{M}$ are canonically isomorphic (up to scalars), which induces a canonical isomorphism between $L$ and $M$ (up to scalars). If $L$ and $M$ are both canonically isomorphic to $\mathbb{C}$ in degree zero, then we can do away with the scalar ambiguity. This becomes particularly bizarre when $M$ admits a filtration such that $\operatorname{gr} M$ is canonically isomorphic to $L$, and this lifts to a filtration on $\mathcal{M}$ such that $\operatorname{gr}\mathcal{M}$ is isomorphic to $\operatorname{gr}\cL$. In this case, you don't really expect $L$ and $M$ to be canonically isomorphic, but rigidity of minimal extension sheaves tells you that they are. In our case we have three modules, $L = I\! H^*_{T^!}(\mathfrak{N}^!)$, $R^{\operatorname{bc}} = \operatorname{Sym}\mathfrak{t}^n/J_{\Delta^{\operatorname{bc}}}$, and $M = H\! P_0(\mathscr{N})$, along with a filtration on $M$ with $\operatorname{gr} M\cong R^{\operatorname{bc}}$ (Lemma \ref{degeneration}). The work of lifting $L$ and $R^{\operatorname{bc}}$ to sheaves $\cL$ and $\mathcal{R}^{\operatorname{bc}}$ on $L_\cH$ and proving that these sheaves are minimal extension sheaves is done in \cite{TP08}. The work of lifting $M$ to a sheaf $\mathcal{M}$ on $L_\cH$ is easy, and we prove that it is a minimal extension sheaf by lifting the filtration on $M$ to one on $\mathcal{M}$ with $\operatorname{gr}\mathcal{M}\cong\cL$. Note that we used exactly this approach in \cite{TP08} with a different module $R$, known as the {\bf Orlik-Terao algebra} of $\cH$. It had already been shown in \cite[Theorem 4]{PS} that $R$ admits a filtration with $\operatorname{gr} R\cong R^{\operatorname{bc}}$, and it was easy to lift that to a filtered sheaf $\cR$ on $L_\cH$ with $\operatorname{gr} \cR\cong\mathcal{R}^{\operatorname{bc}}$. It follows that $\cR$ is a minimal extension sheaf \cite[3.11]{TP08}, and therefore that $R$ is canonically isomorphic to $L$ \cite[4.5]{TP08}. The rest of \cite{TP08} is devoted to showing that the ring structure on $L = I\! H^*_{T^!}(\mathfrak{N}^!)$ induced by this isomorphism can be canonically lifted to define the structure of a ring object on the equivariant $\operatorname{IC}$-sheaf in the equivariant derived category of $\mathfrak{N}^!$ \cite[5.1]{TP08}. \end{remark}
\section{Introduction} The Ooguri-Strominger-Vafa (OSV) conjecture gives a beautiful relation between the partition function of four-dimensional BPS black holes in a type IIA string theory compactified on a Calabi-Yau and the topological A-model string partition function\cite{Ooguri:2004zv}. Consider the BPS black hole partition function, \(Z_{BH}\), in a mixed ensemble given by fixing the magnetic charge, \(p_{\Lambda}\), and summing over the electric charge with electric potential, \(\phi_{\Lambda}\). The OSV conjecture relates the exact microscopic entropy of the black hole, captured by $Z_{BH}$ to the macroscopic entropy, computed in terms of the topological string \begin{equation} Z_{\textrm{BH}}(p^{\Lambda},\phi^{\Lambda}) \sim \vert Z_{\textrm{top}}(X^{\Lambda})\vert ^{2} \label{eqn:osvoriginal} \end{equation} where \(X^{\Lambda} = p^{\Lambda} + \frac{i}{\pi}\phi^{\Lambda}\). This reproduces the Bekenstein-Hawking entropy/area law to the leading order, but otherwise it computes quantum gravitational corrections to it. The entropy on the left is a supersymmetric index. The topological string on the right computes F-terms in the four-dimensional low energy effective action of the IIA string theory on the Calabi-Yau. But from Wald's formula these F-terms are precisely the corrections to the Bekenstein-Hawking black hole entropy. The OSV relation is an example of gauge/gravity duality, where the gauge theory is the theory on the D-branes comprising the black hole. This aspect of the correspondence was emphasized in \cite{Ooguri:2004zv, Vafa:2004qa, Aganagic:2004js ,Ooguri:2005vr, Dijkgraaf:2005bp, Beasley:2006us, Sen:2008yk, Sen:2008vm, Dabholkar:2010uh, Sen:2011ba, Dabholkar:2011ec}.\footnote{The large $N$ dual of a black hole with fixed both magnetic and electric charges is naturally the real version of the topological string, recently studied in \cite{Vafa:2012fi}. Its partition function has both the holomorphic $Z_{top}$ and the anti-holomorphic piece ${\bar Z}_{top}$.} Note that both sides of (\ref{eqn:osvoriginal}) depend only on the K\"ahler moduli and not the complex structure moduli - on the right, this is a well-known property of A-model topological strings, and on the left this is a consequence of computing a well-behaved BPS index. It is natural to ask if there are meaningful ways to generalize the conjecture (\ref{eqn:osvoriginal}). The most general possible black hole partition function would include a chemical potential for angular momentum. This partition function, also known as a spin character, has been extensively studied in the context of motivic wall crossing in four-dimensional \(\mathcal{N}=2\) field theory \cite{KSoriginal, Dimofte:2009tm, Dimofte:2009bv, Cecotti:2009uf, Gaiotto:2010be} and supergravity \cite{Andriyash:2010qv, Andriyash:2010yf}.\footnote{Rotating single-centered black holes in four dimensions cannot be supersymmetric. The black holes that one is studying here are multicentered configurations that carry intrinsic angular momentum in their electromagnetic fields. Note that despite the fact that these are multi-centered, they still correspond to bound states \cite{Denef:2000nb}.} For compact Calabi-Yau manifolds, this spin character will depend sensitively on both the K\"ahler and complex structure of the manifold. On a {\it noncompact} Calabi-Yau, however, the situation improves. We can form a {\it protected spin character} by utilizing the \(SU(2)_{R}\)-symmetry of four-dimensional \(\mathcal{N}=2\) theory \cite{Gaiotto:2010be}. The protected spin character is a genuine index that only depends on the K\"ahler moduli and is constant except at real codimension-one walls of marginal stability. Therefore, this protected spin character gives a well-behaved and computable definition for the refined BPS black hole partition function $Z_{\textrm{ref BH}}(p^{\Lambda},\phi^{\Lambda},y)$, depending on one extra parameter $y$ to keep track of the spin. There is also a natural candidate for what may replace the topological string partition function. The refined topological string is a one-parameter deformation of topological string theory. Just like the refined black-hole partition function, the refined topological string partition function $Z_{\textrm{ref top}}(X^I, y)$ also makes sense only in the non-compact limit, and also utilizes the $SU(2)_R$ symmetry to be defined. The refined topological string is defined as an index in M-theory \cite{Hollowood:2003cv, Iqbal:2007ii}. There are several equivalent ways to compute the index, either by counting spinning M2-branes \cite{Iqbal:2007ii}, or alternatively, as the refined BPS index for D2 and D0-branes bound to a single D6 brane\cite{Iqbal:2003ds , Dijkgraaf:2006um, Dimofte:2009bv}, to name two. Yet another way to compute the index is as the Nekrasov partition function of the five dimensional gauge theory that arises from M-theory at low energies.\footnote{Alternatively, it has a B-model formulation, at least in some cases, in terms of \(\beta\)-deformed matrix models \cite{Dijkgraaf:2009pc, Cheng:2010yw,Aganagic:2011mi}.} Having found a generalization of both $Z_{BH}$ and $Z_{top}$, defined in the same circumstances, and depending on the same set of parameters, it is natural to conjecture that there is in fact a refinement of the OSV conjecture that relates the two: \begin{equation} Z_{\textrm{ref BH}}(p^{\Lambda},\phi^{\Lambda},y) = \vert Z_{\textrm{ref\;top}}(k^{\Lambda},\epsilon_{1},\epsilon_{2})\vert^{2}. \end{equation} Despite the fact that the ingredients for the conjecture fit naturally, one would still like to have a {rationale} for why the conjecture should hold. Because the Calabi-Yau is non-compact, the BPS black holes we are discussing are really BPS particles with large entropy. While one can imagine the system arising by taking a limit where we take the mass of the black-hole to infinity, at the same rate as we take the Planck mass to infinity so that entropy stays finite, the $SU(2)_R$ symmetry we are using makes sense only in the non-compact limit. Correspondingly some of the justifications for the OSV conjecture, for example those based on Wald's formula, may no longer be sound in the refined setting, since the supergravity solution is singular. However, as we mentioned, the original OSV relation is fundamentally a large $N$ duality. For the BPS states that came from a theory on $N$ D-branes, we always get an $SU(N)$ gauge theory describing the particles. In the 't Hooft large $N$ limit of this theory one expects to get a string theory, whether or not there are dynamical gravitons in the theory. Indeed, many famous examples of large $N$ duality are of this kind, see for example \cite{Gopakumar:1998ki}. In particular, the duality should still hold even for the non-compact Calabi-Yau; it is merely difficult to check beyond the protected quantities. Finally, from the perspective of large $N$ duality, once we modify one side of the correspondence, the duality, at least in principle, { fixes} what the other side { has to be}. Furthermore, there is a way to understand directly from the large $N$ duality ${\it why}$ refining the black hole ensemble has to correspond to a refinement of the topological string on the other side. To precisely test the refined OSV conjecture, one would like to compute both sides of the relation and compare explicitly. For the unrefined OSV conjecture, this was done for local, non-compact Calabi-Yau manifolds in \cite{Vafa:2004qa, Aganagic:2004js, Aganagic:2005wn}, and agreement was found. These local Calabi-Yaus can be thought of as the limit of compact Calabi-Yaus in the neighborhood of a shrinking two- (or four-) cycle. In this limit, gravity decouples so we need to be precise about what we mean by the black hole partition function. We must require that the D-branes forming the BPS black hole wrap a cycle that becomes noncompact in this limit so that their entropy remains nonzero. In this limit, the black hole partition function is simply computed by a Witten index on the noncompact brane worldvolume. In the paper, we will run a similar test in the refined case. To this end, we will develop techniques to solve for the refined partition functions on both sides of the conjecture. Remarkably, the refined OSV conjecture passes the tests just as well as the original OSV.\footnote{In subsequent work, \cite{Vafa:2004qa, Dijkgraaf:2005bp, Dabholkar:2005by, Dabholkar:2005dt, Shih:2005he, Aganagic:2006je, Denef:2007vg } several aspects of the conjecture were clarified. For one, while the relation holds to all orders in perturbation theory, the exact relation requires summing over nonperturbative corrections to the macroscopic entropy, taking the form of ``baby universes'' \cite{Dijkgraaf:2005bp, Aganagic:2006je}. Second, at the perturbative level the right side of equation \ref{eqn:osvoriginal} should include a summation over \(\phi \to \phi + 2\pi i n\) so that the periodicities of \(\phi\) match on both sides. Additionally, the right side generically contains an additional measure factor of \(g_{top}^{-2}e^{-K}\), which is natural from the viewpoint of K\"ahler quantization since \(Z_{top}\) transforms as a wavefunction \cite{Verlinde:2004ck, Ooguri:2005vr, Denef:2007vg}. Such a measure factor did not appear in the unrefined local curve examples of \cite{Vafa:2004qa, Aganagic:2004js}. One way to understand this is to observe that these geometries do not have a holomorphic anomaly, which implies that \(Z_{top}\) actually transforms as a function rather than a wavefunction -- therefore the additional measure factors from K\"ahler quantization are absent. In this paper when we study the refined OSV relation on these local geometries, we will again find that no measure factors appear.} This provides strong evidence in support of our conjecture. The paper is organized as follows. In section 2, we review the ingredients of the original OSV conjecture for both compact and non-compact Calabi-Yaus. In section 3, we motivate the refined OSV conjecture first from the perspective of the \(AdS_{2}/CFT_{1}\) correspondence, and second by studying the wall crossing of \(D4\) branes splitting into \(D6\) -\(\overline{D6}\) bound states. In the noncompact setting these arguments are necessarily heuristic, but we believe they capture the correct physics. In section 4, we explain how the refined topological string on Calabi-Yau manifolds of the form \(X =\mathcal{L}_{1}\oplus \mathcal{L}_{2} \to \Sigma_{g}\) can be completely solved by a two-dimensional topological quantum field theory (TQFT). We then use this TQFT to compute refined topological string amplitudes for the above geometries. We also show that our results agree precisely with the five-dimensional Nekrasov partition function of \(U(1)\) gauge theories with \(g\) adjoints that are engineered by these Calabi-Yaus. In section 5, we study the black hole side of the correspondence. Mathematically, the refined partition function for D4/D2/D0-branes computes the \(\chi_{y}\) genus of the relevant instanton moduli spaces. As originally suggested in \cite{Witten:2011zz}, this can be thought of as a categorification of the euler characteristic invariants computed by the \(\mathcal{N}=4\) Vafa-Witten theory. We then specialize to \(D4\) branes wrapping the geometry \(C_{4} = (\mathcal{L}_{1} \to \Sigma_{g})\) and study the refined BPS partition function of bound states with \(D2\) and \(D0\)-branes. We propose that this partition function is computed by a \((q,t)\)-deformation of two-dimensional Yang-Mills, which is closely related to the refined Chern-Simons theory studied in \cite{Aganagic:2011sg, Gadde:2011uv, Aganagic:2012au}. Wrapping branes on the geometry \(C = (\mathcal{O}(-1) \to \mathbb{P}^{1})\), we show that \((q,t)\)-deformed Yang Mills precisely reproduces a mathematical result of Yoshioka and Nakajima for the \(\chi_{y}\) genus of instanton moduli space \cite{Yoshioka:1996pd, Nakajima:2003uh}. In section 6 we connect the black hole and topological string perspectives by studying the large N limit of the \((q,t)\)-deformed Yang Mills theory. We find that the theory factorizes to all orders in \(1/N\) into two copies of the refined topological string partition function. This gives a nontrivial check of our refined OSV conjecture. Finally, in section 7 we explain an alternative way to compute the refined black hole partition function on \(\mathcal{O}(-1) \to \mathbb{P}^{1}\). The refined bound states are counted by using the semi-primitive refined wall-crossing formula \cite{Dimofte:2009bv}, thus giving a refined extension of the techniques used in \cite{Nishinaka:2010qk, Nishinaka:2010fh}. \section{The OSV Conjecture: Unrefined and Refined \label{sec:osv}} We start by reviewing the remarkable conjecture of Ooguri, Strominger, and Vafa (OSV) connecting four-dimensional BPS black holes with topological strings \cite{Ooguri:2004zv}. Consider IIA string theory compactified on a Calabi-Yau, \(X\), with four-dimensional black holes arising from D-branes wrapping holomorphic cycles in \(X\). In terms of four-dimensional gauge fields, the \(D0\) and \(D2\) branes are electrically charged while the \(D4\) and \(D6\) branes are magnetically charged. The object of interest for the OSV conjecture is the mixed black hole partition function given by fixing the magnetic charges and summing over electric charges with chemical potentials, \begin{equation} Z_{BH}(P_{6},P_{4};\phi_{2},\phi_{0}) = \sum_{Q_{2},Q_{0}}\Omega(P_{6},P_{4},Q_{2},Q_{0})e^{-\phi_{2}Q_{2}-\phi_{0}Q_{0}} \end{equation} where we have denoted \(D6\) charges by \(P_{6}\), \(D4\) charges by \(P_{4}\), \(D2\) charges by \(Q_{2}\), and \(D0\) charges by \(Q_{0}\). Here \(\phi_{2}\) and \(\phi_{0}\) are chemical potentials associated to the electrically charged D-branes. \(\Omega(P,Q)\) is computed by the Witten index in the corresponding charge sector, \begin{equation} \Omega(P_{6},P_{4},Q_{2},Q_{0}) = \textrm{Tr}_{\mathcal{H}_{P,Q}}(-1)^{F} \end{equation} and only receives contributions from BPS black holes. The OSV conjecture states that this mixed black hole partition function is equal to the square of the A-model topological string partition function on \(X\), \begin{equation} Z_{BH}(P_{6},P_{4};\phi_{2},\phi_{0}) = \vert Z_{top}(g_{top},k)\vert^{2} \end{equation} where \(k\) is the complexified K\"ahler form on \(X\). The projective coordinates on moduli space are given by, \begin{equation} X_{I} = P_{I} + i \frac{\phi_{I}}{\pi} \end{equation} This implies that the string coupling constant and K\"ahler moduli are determined by the magnetic charges and electric potentials, \begin{eqnarray} g_{top} & = & \frac{4\pi i}{X_{0}} = \frac{4\pi i}{P_{6}+i\frac{\phi_{0}}{\pi}} \\ k_{I} & = & 2\pi i \frac{X_{I}}{X_{0}} = \frac{1}{2}g_{top}\Big(P_{4,I}+\frac{i\phi_{2,I}}{\pi}\Big) \end{eqnarray} Here the real part of \(X_{I}\) is fixed by the attractor mechanism which determines the near-horizon Calabi-Yau moduli in terms of the black hole charge. Both sides of the relation should be considered as expansions in \(1/Q\) where \(Q\) is the total graviphoton charge of the black hole. From the change of variables, we have a perturbative expansion in \(g_{top}\) if either \(P_{6}\) or \(\phi_{0}\) is large. We will usually set \(P_{6}=0\) so it is natural to take both \(\phi_{0}\) and \(P_{4}\) to infinity is such a way that \(g_{top}\) becomes small and the K\"ahler form remains constant. One way to further understand the OSV relation is by inverting it, \begin{equation} \Omega(P_{6},P_{4},Q_{2},Q_{0}) = \int d\phi_{0}d\phi_{2} e^{Q_{2}\phi_{2}+Q_{0}\phi_{0}} \vert Z_{top}\vert^{2} \end{equation} so that black hole degeneracies are formally computed by the topological string. It is known that because the topological string partition function obeys the holomorphic anomaly equations \cite{Bershadsky:1993cx}, it transforms as a wavefunction under changes of polarization on the Calabi-Yau moduli space \cite{Witten:1993ed}. Thus, from quantum mechanics the appearance of \(\vert Z_{top} \vert^{2}\) is very natural. In fact, this interpretation implies that \(\Omega(P,Q)\) is the Wigner quasi-probability function on \((P,Q)\) phase space \cite{Ooguri:2004zv}. The original OSV conjecture focused on the case of compact Calabi-Yau manifolds, so that the wrapped D-branes correspond to black holes in four-dimensional \(\mathcal{N}=2\) supergravity. However, as explored in \cite{Vafa:2004qa, Aganagic:2004js, Aganagic:2005dh, Dijkgraaf:2005bp, Aganagic:2005wn, Aganagic:2006je}, it is interesting to study the OSV conjecture for local Calabi-Yau manifolds which can be thought of as the decompactification limit of the compact case. In this limit gravity decouples which means that the four dimensional planck mass goes to infinity. Since the Bekenstein-Hawking entropy of a black hole is proportional to \(M_{BH}^{2}/M_{Pl}^{2}\), to obtain a finite entropy in this limit we should also take \(M_{BH}\to\infty\). This can be accomplished simply by wrapping \(D4\) or \(D6\)-branes on cycles that become non-compact in this limit. The precise OSV relation remains the same in this limit, except that now the black hole partition function is naturally computed by a partition function on the worldvolume of the noncompact \(D4\) or \(D6\)-branes. The advantage of taking this limit is that both sides of the OSV relation are exactly solvable, leading to a highly nontrivial check of the conjecture. The conjecture has been tested perturbatively to all orders in \cite{Vafa:2004qa, Aganagic:2004js, Aganagic:2005wn}, and non-perturbative corrections in the form of baby universes have been computed in \cite{Dijkgraaf:2005bp, Aganagic:2006je}. \subsection{Refining the Conjecture} Now that we have reviewed the OSV conjecture, a natural question to ask is whether the black hole degeneracy computed by \(\Omega(P,Q)\) is the most general index that counts four-dimensional BPS black holes. In fact, we could include information about spin by replacing the Witten index, \begin{equation} \textrm{Tr}_{\mathcal{H}_{BPS}}(-1)^{F} \end{equation} by the spin character, \begin{equation} \textrm{Tr}_{\mathcal{H}_{BPS}}(-1)^{F}\exp(-2\gamma J_{3}) \end{equation} where \(J_{3}\) is the three-dimensional generator of rotations and \(\gamma\) is the conjugate chemical potential.\footnote{Rotating single-centered black holes in four dimensions cannot be supersymmetric. The black holes that we are studying here are multicentered configurations that carry intrinsic angular momentum in their electromagnetic fields. Note that despite the fact that these are multi-centered, they still typically correspond to bound states \cite{Denef:2000nb}.} These spin-dependent BPS traces and their wall-crossing behavior have been studied extensively in the context of \(\mathcal{N}=2\) field theory \cite{KSoriginal, Dimofte:2009bv, Dimofte:2009tm, Cecotti:2009uf, Gaiotto:2010be} and supergravity \cite{Andriyash:2010qv, Andriyash:2010yf}. However, this trace has the drawback of not being an index, which means that it will be sensitive to both the complex and K\"ahler moduli. If we consider a local Calabi-Yau manifold by taking the gravity decoupling limit, there is a preserved \(SU(2)\) R-symmetry that appears. As explained in \cite{Gaiotto:2010be}, this can be used to form a protected spin character that is a genuine index, and only depends on the K\"ahler moduli through wall-crossing, \begin{equation} \textrm{Tr}_{\mathcal{H}_{P,Q}}(-1)^{2J_{3}}e^{-2\gamma(J_{3}-R)} = \sum_{J_{3},R}\Omega(P,Q;J_{3},R)e^{-2\gamma(J_{3}-R)} \end{equation} Now we can form the mixed ensemble of black holes counted with spin where, as in the ordinary case, we fix the magnetic charge and sum over the electric charge, \begin{equation} Z_{\textrm{ref BH}}(P_{6},P_{4};\phi_{2},\phi_{0};\gamma) = \sum_{Q_{2},Q_{0},J_{3},R}\Omega(P_{6},P_{4},Q_{2},Q_{0};J_{3},R)e^{-2\gamma (J_{3}-R)-\phi_{2}Q_{2}-\phi_{0}Q_{0}} \end{equation} We will refer to this as the refined black hole partition function. We would like to know whether there exists a generalization of the topological string whose square is equal to \(Z_{\textrm{ref BH}}\). A natural candidate for this one-parameter deformation is well-known, and is given by the refined topological string. Recall the definition of the refined topological string as the index of M-theory, depending on Kahler moduli $k$ and two additional parameters $\epsilon_1$ and ${\epsilon_2}$. The refined topological string partition function \cite{Hollowood:2003cv, Iqbal:2007ii, Aganagic:2011mi} on a Calabi-Yau $X$ is given by computing the index of M-theory on the geometry, \begin{equation} (X\times TN \times S^{1})_{\epsilon_1,\epsilon_2} \end{equation} where \(TN\) denotes the Taub-NUT spacetime, and upon going around the \(S^{1}\) the Taub-NUT is twisted by, \begin{eqnarray} (z_{1}\, , \, z_2 ) \qquad \to \qquad (q\, z_{1} , \,t^{-1}\,z_{2} ) \end{eqnarray} where $$ q=e^{-\epsilon_{1}}\qquad t=e^{-\epsilon_{2}}. $$ In addition, we must include an R-symmetry twist to preserve supersymmetry. This twist is implemented by a geometric Killing vector on the non-compact Calabi-Yau, \(X\). Note that, in our notation, the unrefined limit is $\epsilon_1=\epsilon_2$, unlike in much of the literature, where one typically defines $\epsilon_2$ with a different sign. The partition function of M-theory in this geometry is computing the index of the resulting theory on $TN\times S^1$, \begin{equation} Z_{ref\; top}(\epsilon_1,\epsilon_2;k) = \textrm{Tr}(-1)^{2S_{1}+2{S_2}}q^{S_{1}-{R}}t^{{R}-S_{2}}e^{- k_I Q^I_2} \label{eqn:refindex} \end{equation} where \(S_{1}\) and \(S_{2}\) are the spins in the \(z_{1}\) and \(z_{2}\) directions, respectively, and \({R}\) is the R-charge of the state. We have schematically indicated that the partition function depends on the Kahler moduli $k$, via the M2 brane contributions to the index. Note that although the trace is over all states, only BPS states will make a contribution. The index can be computed in several different ways: by counting spinning M2-branes \cite{Hollowood:2003cv, Iqbal:2007ii} or as the Omega-deformed instanton partition function \cite{Nekrasov:2002qd}. In analogy with the unrefined case \cite{Iqbal:2003ds}, the refined topological string can also be written in terms of the refined Donaldson-Thomas invariants that compute the BPS protected spin character for D2 and D0-branes bound to a single D6 brane\cite{Dimofte:2009bv}. Given that the refined topological string also counts BPS particles with spin and only depends on the K\"ahler moduli of \(X\), it is should be related to the refined black hole partition function. The index is related to the ordinary topological string partition function is we set $\epsilon_1=\epsilon_2 = g_s$, and reduce on the thermal $S^1$ to IIA. Then, we get IIA string theory on $X\times TN$, whose partition function is the same as the ordinary topological string partition function, where $g_s$ is the topological string coupling constant \cite{Dijkgraaf:2006um}. In particular, M2 branes wrapping holomorphic curves in $X$ and the thermal $S^1$ become the worldsheet instantons of the topological string. For $\epsilon_1\neq \epsilon_2$, the theory has no known worldsheet formulation, at the moment. There is yet another way to view the partition function (\ref{eqn:refindex}), which will be useful for us. This corresponds to the TST dual formulation, where instead, we go down to IIA string theory on the $S^1$ in the Taub-Nut space \cite{Dijkgraaf:2006um}. This turns the Taub-Nut space into a single D6 brane wrapping $X\times S^1$. In this case, the refined topological string partition function has the interpretation as the refined spin character, counting the bound states of the D6 brane on $X$ with D0 and D2 branes. In terms of the $SO(4)=SU(2)_{\ell }\times SU(2)_r$ rotation symmetry of the Taub-Nut space, the D0 brane charge $Q_0$ is the $2J_3^{\ell}$ component of the $SU(2)_{\ell}$ spin in M-theory, the $SO(3)=SU(2)$ rotation symmetry in IIA is identified, under the dimensional reduction, with the $SU(2)_r$ symmetry in M-theory, while the $SU(2)_R$ R-symmetry is manifestly the same in both IIA and M-theory. In particular, the refinement is associated with the diagonal $SU(2)_d \subset SU(2)_r\times SU(2)_R$ spin\cite{Nekrasov:2003rj}. This allows us to rewrite \ref{eqn:refindex} as \begin{equation} Z_{ref\; top}(\epsilon_1,\epsilon_2;k) = \textrm{Tr}(-1)^{2J_3}\; e^{\frac{\epsilon_1 +\epsilon_2}{2}Q_0}\; e^{\frac{\epsilon_1 -\epsilon_2}{2}(2J_3 - 2R)}e^{k_I Q^I_2} \label{eqn:refindexD}. \end{equation} In writing this, we used the fact that $2J_3^{\ell} = S_1-S_2$, $2J_3^{r}=S_1+S_2$, which is obvious from the way the $SU(2)_{\ell}\times SU(2)_r$ acts on the coordinates $z_1,z_2$ of the Taub-Nut space, and furthermore, as we just reviewed, that $Q_0=2J^{\ell}_3$ and $2J_3=2J_3^r$. This leads us to propose the refined OSV conjecture relating the protected spin character of black holes to the refined topological string\footnote{For an alternate proposal relating the Nekrasov partition function to non-supersymmetric extremal black holes, see \cite{Saraikin:2007jc}.}, \begin{equation} Z_{\textrm{ref BH}}(P_{6},P_{4};\phi_{2},\phi_{0};\gamma) = \vert Z_{\textrm{ref top}}(\epsilon_{1},\epsilon_{2},k)\vert^{2} \label{eqn:refosv} \end{equation} To complete the conjecture, we propose that the variables are related by, \begin{eqnarray} k_{I} & = & \frac{2\pi i (\frac{i\phi_{2,I}}{\pi} + \beta P_{4,I})}{\frac{i\phi_{0}}{\pi} + \beta P_{6}} \nonumber \\ \epsilon_{1} & = & \frac{4\pi i C}{i\frac{\phi_{0}}{\pi}+ \beta P_{6}} \label{eqn:osvchange} \\ \epsilon_{2} & = & \frac{4\pi i C \beta}{i\frac{\phi_{0}}{\pi}+ \beta P_{6}} \nonumber \end{eqnarray} where we have defined the variable \(\beta \equiv 1 - \frac{\gamma}{2\pi i}\), and we have included an additional constant \(C\).\footnote{As explained in \cite{Ooguri:2004zv}, an arbitrary constant $C$ is needed in the compact case due to the fact that $X_I$ are not functions on the moduli space, but sections of a line bundle. In the non-compact case, which we study, this degree of freedom is fixed. In the unrefined case, it is typically set to $1$. Our choice of the refined value is such that it reduces to $1$ when we set $\epsilon_{1,2}$ to be equal.} The most natural choice for \(C\), as we explain in section \ref{sec:motiv} is \begin{equation} C = \frac{2\epsilon_{1}}{\epsilon_{1} + \epsilon_{2}} \end{equation} so that when \(P_{6}=0\) we have, \begin{equation} \phi_{0} = \frac{8\pi^{2}}{\epsilon_{1} + \epsilon_{2}} \end{equation} for the D0 brane chemical potential, and moreover $$ {\gamma} ={2\pi i} \frac{\epsilon_{1} - \epsilon_{2}}{\epsilon_{1} + \epsilon_{2}} $$ for the spin chemical potential. In the next section we will motivate the conjecture, and explain the origin of the change of variables. Note that, in the specific example that we study in sections \ref{sec:tqft}-\ref{sec:factor}, we will find that one gets a slightly different effective value of \(C\), for the reason we will explain (having to do with a shift in the zero of the spin for the D0 branes). \section{Motivating the Refined Conjecture \label{sec:motiv}} As explained in \cite{Ooguri:2004zv, Vafa:2004qa, Ooguri:2005vr, Dijkgraaf:2005bp, Beasley:2006us, Sen:2008yk, Sen:2008vm, Dabholkar:2010uh, Sen:2011ba, Dabholkar:2011ec}, the OSV conjecture is an instance of large $N$ duality. In this case, the gauge theory is the $SU(N)$ gauge theory on the $N$ D-branes wrapping cycles of the Calabi-Yau manifold $X$ and comprising the black holes. The large $N$ dual of this theory is a string theory in the back-reacted geometry. In the full physical string theory, the near-horizon geometry of a BPS black hole in four dimensions takes the form \begin{equation} AdS_{2} \times S^{2} \times X. \end{equation} The OSV conjecture deals with supersymmetric sub-sectors of the theory. On the black-hole side, we consider the Witten index of the theory on $N$ D-branes; this is typically a partition function of a topological $SU(N)$ gauge theory in one dimension less. On the large $N$ dual side, the partition function ends up depending only on F-terms in the low energy effective action, which are captured by the topological string partition function. The OSV conjecture can also be thought of as a consequence of large $N$ duality in the topological setting alone. The 't Hooft large $N$ duality, relating a $SU(N)$ gauge theory to a string theory is a very general phenomenon. It should encompass any $SU(N)$ gauge theory, including topological ones, and requires a string theory on the dual side, though not one containing dynamical gravity. In particular, in \cite{Vafa:2004qa, Aganagic:2004js} it was shown that OSV conjecture holds even for non-compact Calabi-Yau manifolds. In the physical version of the theories studied there, the Planck mass is infinite, and the the black holes horizon will have zero area, making it difficult to study the large $N$ duality in the full physical theory. In the refined context, we have to take the Calabi-Yau to be non-compact, since the R-symmetry which is necessary to compute the protected index exists only in that case. However, the theory on $N$ D-branes is still a $SU(N)$ gauge theory, with large entropy at large $N$. $Z_{\textrm{ref BH}}$ is simply a one parameter deformation of the ordinary black-hole partition function $Z_{\textrm{BH}}$. The dual description of the theory at large $N$ has to be a string theory, on general grounds, and moreover, a suitable one-parameter deformation of the topological string theory. Now, we will explain why this one-parameter deformation is the refined topological string. The statement of the refined OSV conjecture is that we can refine both sides of the OSV duality, by keeping track of the ${J_3-R}$ charge. Why this should be true is most tranparent in yet another way to understand the OSV, namely, using wall crossing \cite{Denef:2007vg}. In this case, we do not take the near-horizon limit but instead we study general D4/D2/D0-brane bound states. We can then perform \(TST\)-duality on this partition function so that it is dominated by ``polar'' states. Generically, these polar states can be made to decay by varying the background K\"ahler parameters. Along a real co-dimension one wall, the state will decay into a D6/D4/D2/D0 state and a \(\overline{D6}\)/D4/D2/D0-state. By a chain of dualities (lifting to M-theory, then reducing on a different circle), the \(AdS_{2} \times S^{2} \times X\) geometry (see \cite{Beasley:2006us, Denef:2007vg} for details) can be related to IIA string theory on $X$, with a \(D6-\overline{D6}\) pair. Further, from the primitive wall-crossing formula we know that the degeneracies will factorize, \begin{equation} \Omega(D4+\ldots) \sim \Omega(D6+\ldots)\Omega(\overline{D6}+\ldots) \end{equation} Now the key observation is that the degeneracies of D6/D4/D2/D0 brane bound states are precisely computed by Donaldson-Thomas invariants, which are further identified with the topological string. On the topological string side, the S-duality used is precisely what relates the D6 brane partition function to the topological string \cite{Nekrasov:2004js}, as we reviewed in the previous section. Therefore, we can identify the D6-brane bound states with \(Z_{top}\) and the \(\overline{D6}\)-brane bound states with \(\overline{Z}_{top}\). Therefore, the semiprimitive wall-crossing formula gives precisely the factorization expected from OSV, \begin{equation} Z_{BH} \sim \vert Z_{top} \vert^{2} \end{equation} In the refined setting, the argument goes through in precisely the same way as in the unrefined case; we simply replace, on both sides of the duality, the Witten index of the D4 brane and the D6 branes by the protected spin character. Then we simply use the refined primitive wall-crossing formula which also factorizes. On the black hole side, the protected spin character of the D4 branes is the refined black hole partition function $Z_{\textrm{ref BH}}$, and on the topological string side, the protected spin character of the D6 branes is the topological string partition function -- moreover, we get both $Z_{\textrm{top ref}}$ and ${\overline Z}_{\textrm{top \;ref}}$ from the D6 branes and the $\overline{D6}$ branes, and thus \begin{equation} Z_{\textrm{ref BH}} \sim \vert Z_{\textrm{ref top}} \vert^{2}. \end{equation} This argument makes it obvious that the OSV conjecture should hold in the refined setting, as we conjectured. The one subtlety in this argument is that on a noncompact Calabi-Yau geometry, there is actually no place in moduli space where the \(D4\)-branes can be made to split into \(D6-\overline{D6}\) constituents, since D6 and \(\overline{D6}\) branes will always have opposite central charges because of the noncompactness of the Calabi-Yau. However, this issue was already present in the unrefined case, where it did not affect the validity of the conjecture, as was shown in \cite{Vafa:2004qa,Aganagic:2004js}. Thus, there is no reason to think it would affect our refined conjecture either. Thus, we believe that this \(D6\)-\(\overline{D6}\) decomposition captures the correct physics of D4/D2/D0-brane bound states in both the unrefined and the refined case, and this moreover leads to a refined OSV formula. In the rest of this section, we will provide further support for the conjecture, and explain the identification of the parameters we gave previously. \subsection{Refined OSV and The Wave Function on the Moduli Space} The refined topological string partition function is a wave function on the moduli space, just like in the ordinary topological string case \cite{Krefl:2010fm, Krefl:2010jb, Huang:2011qx, Aganagic:2011mi}. The quantum mechanics on the moduli space played a central role in understanding the original conjecture, and the same is true in the refined case. In this respect, there only two differences between the refined and the unrefined topological string thing: for one, the effective value of the Plank's constant of the theory $g_s^2$, becomes $\epsilon_1\epsilon_2$ (recall that, in the unrefined case, $\epsilon_1$ and $\epsilon_2$ coincide), $$ g_s^2\qquad \rightarrow \qquad g_s^2=\epsilon_1 \epsilon_2. $$ Secondly, the wave function that the topological string partition function computes changes: in the refined case, this wave function depends on the additional parameter $\beta =\epsilon_2/\epsilon_1$. For this discussion of the quantum mechanics, it is useful to switch to the mirror perspective and study the refined B-model \cite{Aganagic:2011mi}. The refined B-model only depends on the complex structure moduli space, which can be parametrized by the holomorphic three form \(\Omega \in H_{3}(X)\). We can choose a symplectic basis for \(H_{3}(X)\) such that \(A_{I}\cap B^{J} = \delta^{J}_{I}\), and define coordinates, \begin{equation} X_{I} = \int_{A_{I}}\Omega, \qquad\qquad F^{J} = \int_{B^{J}}\Omega \end{equation} From special geometry, we know that classically these variables are not independent and that there exists a prepotential, \(F^{(0)}\), such that, \begin{equation} F^{J} = \frac{\partial F^{(0)}}{\partial X_{J}} \end{equation} But now it is important to recognize that this prepotential is the genus zero contribution of the refined topological string, \begin{equation} Z_{\textrm{ref top}} = \exp\Big(\frac{1}{\epsilon_{1}\epsilon_{2}}F\Big) = \exp\Bigg(\frac{1}{\epsilon_{1}\epsilon_{2}}F^{(0)} + \ldots\Bigg) \end{equation} Therefore in the full quantum theory, we can represent \(F^{J}\) as the operator \(F^{J} = \epsilon_{1}\epsilon_{2}\frac{\partial}{\partial X_{J}}\) and this leads to the commutation relations, \begin{equation} \lbrack F^{J}, X_{I} \rbrack = \epsilon_{1}\epsilon_{2} \delta^{J}_{I} \end{equation} We could have applied this same reasoning to the conjugated theory, \(\overline{Z}_{\textrm{ref top}}\) which gives, \begin{equation} \lbrack \overline{F}^{J}, \overline{X}_{I} \rbrack = \epsilon_{1}\epsilon_{2} \delta^{J}_{I} \end{equation} and finally all of the barred variables commute with all of the unbarred variables. Note that, in this case, because the Calabi-Yau is non-compact, the moduli space is always governed by the rigid special geometry of the ${\cal N}=2$ field theory, rather than the local special geometry of ${\cal N}=2$ supergravity. Now consider formally inverting the refined OSV relation, \begin{equation}\label{eq-overlap} \Omega(P_{6},P_{4},Q_{2},Q_{0};\gamma) = \int d\phi_{0}d\phi_{2} e^{Q_{2}\phi_{2}+Q_{0}\phi_{0}} \vert Z_{\textrm{ref top}}\vert^{2} \end{equation} where \begin{equation} \Omega(P_{6},P_{4},Q_{2},Q_{0};\gamma) = \sum_{J_{3},R}\Omega(P_{6},P_{4},Q_{2},Q_{0};J_{3},R)e^{-2\gamma(J_{3}-R)}. \end{equation} We say that this is a formal inversion, since in the non-compact case $\phi_0$ is always just a parameter, so in particular, it does not really make sense integrating over it. This aside, note that for the relation such as (\ref{eq-overlap}) to make sense, it {\it has to be the case} that $Z_{\textrm{ref top}}$ is indeed a wave function on the moduli space. This is because while the left hand side is independent of the choice of polarization, i.e. the choice of basis of $A$- and $B$-cycles, for an arbitrary function on the moduli space, the right hand side {\it would} depend on such a choice, and the conjecture would not have a chance to hold. Because $Z_{\textrm{ref top}}$ is a wave function, while all the terms on the right hand side depend on the choice of polarization, the integral does not depend on such a choice. More precisely, for this to hold, one has to have the following commutation relations. We follow the reasoning in \cite{Ooguri:2004zv}, and formally introduce magnetic potentials, \(\chi^{I}\) in addition to the electric potential that we have already used. In the refined black hole partition function, we cannot specify both the electric charge and the electric potential at the same time, so they must have nontrivial commutation relations. Similarly, we require that the new magnetic potentials are conjugate to the magnetic charges. Therefore, we find \begin{eqnarray} \lbrack \phi_{I}, Q^{J} \rbrack & = & \lbrack P_{I}, \chi^{J} \rbrack = \frac{i\pi}{2}\delta^{J}_{I} \nonumber \\ \lbrack \phi_{I}, P_{J} \rbrack & = & \lbrack \chi^{I}, Q^{J} \rbrack = 0 \\ \lbrack Q^{I}, P_{J} \rbrack & = & \lbrack \chi^{I}, \phi_{J} \rbrack = 0 \nonumber \end{eqnarray} where for convenience we have included an extra normalization factor above. For the right hand side of \ref{eq-overlap} to be invariant under symplectic transformations one needs the black hole commutation relations and the topological string commutation relations, to be consistent. This requires, \begin{eqnarray} X_{I} & = & C'\epsilon_{2} P_{I} + i\frac{\epsilon_{1}}{C'}\frac{\phi_{I}}{\pi} \label{eqn:xf} \\ F^{I} & = & C'\epsilon_{2} Q^{I} + i\frac{\epsilon_{1}}{C'}\frac{\chi^{I}}{\pi} \nonumber \end{eqnarray} for some arbitrary constant, \(C'\). Note that is in prefect agreement with the refined OSV change of variables in equation \ref{eqn:osvchange} upon fixing the constant to \(C'=1\). In fact, these relations were our main motivation for the change of variables we proposed in section 3, as a part of our conjecture. Notice that \(\Omega(P,Q;\gamma)\) still has the interpretation as a Wigner quasi-probability distribution on phase space, just as it did in \cite{Ooguri:2004zv}, but now it depends on the additional auxillary parameter, \(\gamma\). To understand the identification of $\epsilon_1$ and $\epsilon_2$ with the D0 brane chemical potential $\phi_0$ and the spin fugacity $\gamma$, we can use the wall crossing derivation, which forces the identification of parameters. The only subtlety is that, to relate the refined topological string to the black-hole ensemble, we need to perform the TST duality. The TST duality relates this to the chemical potentials before and after as follows: \begin{equation}\label{Sduality} \phi_{0}\;\; \to \;\;\phi_{0}'=\frac{4\pi^{2}}{\phi_{0}}, \qquad \phi_{2} \;\; \to\;\; \phi_{2}'=2\pi i \frac{\phi_{2}}{\phi_{0}}, \qquad \gamma \;\;\to \;\;\gamma'=2\pi i \frac{\gamma}{\phi_{0}} \end{equation} The derivation of this is presented in appendix \ref{sec:sduality}. As we reviewed in the previous section, the chemical potential for the D0 branes bound to the D6 brane is \begin{equation} \phi'_{0} = \frac{\epsilon_{1} + \epsilon_{2}}{2}, \end{equation} and the spin is captured by \begin{equation} \gamma' = \frac{\epsilon_{1} - \epsilon_{2}}{2}, \end{equation} For this to be consistent with TST duality, the chemical potentials in the black hole ensemble need to be \begin{equation} \phi_{0} = \frac{4\pi^{2}}{\epsilon_{1} + \epsilon_{2}}, \end{equation} for the D0 brane charge, and moreover the spin needs to be captured by \begin{equation} \gamma = {2 \pi i}\frac{\epsilon_{1} - \epsilon_{2}}{\epsilon_{1} + \epsilon_{2}} \end{equation} just as we gave in the previous section. In particular, $\gamma/2\pi i = 1-\beta$. The rest of this paper is devoted to testing our conjecture. There is a class of geometries where both sides of duality are computable explicitly. These correspond to Calabi-Yau manifolds that are complex line bundles over a Riemann surface. After developing the necessary tools to precisely compute both sides of the refined OSV formula, we show that the refined OSV conjecture holds true perturbatively to all orders for these geometries. \section{Refined Topological String on ${\mathcal L}_1\oplus {\mathcal L}_1 \rightarrow \Sigma $ \label{sec:tqft}} For some simple Calabi-Yau manifolds, the refined topological string partititon function is exactly computable by cutting the Calabi-Yau into simple pieces, and sewing them back together. The open-string version of this index was computed explicitly on simple geometries in \cite{Aganagic:2011sg} and used to solve the refined Chern-Simons theory completely. We will follow a similar approach here in the closed string case, for Calabi-Yau manifolds of the form $$ {\mathcal L}_1\oplus {\mathcal L}_1 \rightarrow \Sigma $$ To obtain a Calabi-Yau manifold, the degrees of the line bundles must satisfy the property, \begin{equation} \textrm{deg}(\mathcal{L}_{1}) + \textrm{deg}(\mathcal{L}_{2}) = -\chi(\Sigma) = 2g-2 \end{equation} The key idea is to chop up our geometries by introducing stacks of infinitely many M5 brane/anti-brane pairs wrapping Lagrangian three-cycles as in the original topological vertex \cite{Aganagic:2003db}.\footnote{In the refined setting, we must choose whether to wrap these M5 branes on the \(z_{1}\) or \(z_{2}\) plane of the Taub-NUT space. This gives two types of refined A-branes, which can be denoted as \(q\)-branes and \(t\)-branes. At each boundary of our geometry we can place either type of brane, leading to different choices of basis for each Hilbert space. In this paper we will not need this rich structure, and we will implicitly place \(q\)-branes at each puncture. We refer the reader to \cite{AS2, ASVertex} for details on general \(q\)/\(t\)-brane amplitudes.} Then the computation of the refined index on these chopped geometries reduces to counting M2 branes ending on these M5 branes. In this paper, we simply explain the structure of the TQFT, and refer the reader to \cite{ASVertex} for the details of computing these amplitudes by counting M2-brane contributions. \subsection{A TQFT for the Refined Topological String} The basic building blocks of the TQFT are given by the annulus (A), cap (C), and pant (P) geometries. Since degrees of bundles and euler characteristics add upon gluing, this gives a way of building up more complicated bundles over Riemann surfaces. We start by considering the simplest geometry, which is the annulus (shown in Figure \ref{fig:tqftpieces}) with two trivial complex line bundles over it given by \(A^{(0,0)} = \mathbb{C}^{*} \times \mathbb{C}^{2}\). \begin{eqnarray} Z_{\textrm{ref top}}(A^{(0,0)}) & = & \sum_{R}\frac{1}{g_{R}(q,t)}M_{R}(U;q,t)M_{R}(V;q,t) \end{eqnarray} Here we are summing over all \(U(\infty)\) representations \(R\), and \(M_{R}(U;q,t)\) is the associated Macdonald polynomial which reduces on setting \(q=t\) to the simpler \(\textrm{Tr}_{R}(U)\). The Macdonald metric, \(g_{R}\) computes the inner product of a Macdonald polynomial with itself and is given by,\footnote{Note we are including additional (q/t) factors in our definition of $g_{R}$ and \(M_{R}(U)\) compared to the standard definitions in \cite{macdonald_hall}. The advantage of these factors is that they restore the symmetry, \(g_{R}(q,t) = g_{R}(q^{-1},t^{-1})\) (see also \cite{Awata:2008ed} for a similar shift). We refer the reader to Appendix \ref{sec:macs} for more details on our Macdonald polynomial conventions.} \begin{eqnarray} g_{R}(q,t) & = & (t/q)^{\vert R \vert/2}\prod_{(i,j)\in R}\frac{1-t^{R^{T}_{j}-i}q^{R_{i}-j+1}}{1-t^{R^{T}_{j}-i+1}q^{R_{i}-j}} \label{eqn:uinftymetric} \\ & = & \prod_{(i,j)\in R}\frac{t^{\frac{R^{T}_{j}-i}{2}}q^{\frac{R_{i}-j+1}{2}}-t^{-\frac{R^{T}_{j}-i}{2}}q^{-\frac{R_{i}-j+1}{2}}}{t^{\frac{R^{T}_{j}-i+1}{2}}q^{\frac{R_{i}-j}{2}} - t^{-\frac{R^{T}_{j}-i+1}{2}}q^{-\frac{R_{i}-j}{2}}} \label{eqn:infmetric} \end{eqnarray} Since we are working with \(U(\infty)\) representations, this metric is the \(N\to\infty\) limit of the ordinary \(SU(N)\) Macdonald metric. We refer the reader to Appendix \ref{sec:macs} for our Macdonald polynomial conventions. Having discussed the simplest geometry, we should now explain how building blocks are glued together. Recall that ordinarily when we want to glue two boundaries together, we should set their holonomies to be equal except that the boundaries should have opposite orientation. This orientation reversal simply flips one of the holonomies from \(U\) to \(U^{-1}\). Finally, to glue together the boundaries we must integrate over the Hilbert space at the boundaries. This is also true in the refined setting, except that the integration measure is deformed to the one natural for Macdonald polynomials. If we denote the eigenvalues of \(U\) by \(e^{u_{i}}\), then the Macdonald measure is given by, \begin{equation} \Delta(U;q,t) = \prod_{m=0}^{\infty}\prod_{i\neq j}\frac{\Big(1-q^{m}e^{u_{i}-u_{j}}\Big)}{\Big(1-tq^{m}e^{u_{i}-u_{j}}\Big)} \end{equation} Then gluing two boundaries gives, \begin{equation} \int du_{i}\Delta(U;q,t)M_{R_{1}}(U)M_{R_{2}}(U^{-1}) = g_{R_{1}}(q,t)\delta_{R_{1}R_{2}} \end{equation} where \(g_{R}\) is the Macdonald metric for infinitely many variables introduced above, which is to be contrasted with the finite N Macdonald metric which we will define below in equation \ref{eqn:metricbeta}. Thus, the \(M_{R}\) give an orthogonal but not orthonormal basis for the boundary Hilbert space. Although we could remove the explicit metric factors \(g_{R}\) by choosing a different normalization for \(M_{R}\), it will actually be more convenient in this paper to keep them. As a simple consistency check, gluing two annuli with trivial bundles should give back the original annulus amplitude. But this is clearly true, since the two annuli contribute a total factor of \(g_{R}^{-2}\) while the gluing process contributes a factor of \(g_{R}\) so that the resulting amplitude is equal to the original annulus amplitude. Now that we have explained gluing, we also want to know how to introduce nontrivial bundles. Note that since \(\chi(A)=0\), any choice of line bundles over the annulus must satisfy \(\textrm{deg}(\mathcal{L}_{1}) = -\textrm{deg}(\mathcal{L}_{2})\). The simplest nontrivial choice is the geometry \(A^{(1,-1)}\). This geometry can be alternatively understood as implementing a change in framing, which has been studied for the refined topological vertex in \cite{Iqbal:2007ii}. The resulting amplitude is given by, \begin{equation} Z_{\textrm{ref top}}(A^{(1,-1)}) = \sum_{R}\frac{1}{g_{R}(q,t)}q^{\frac{1}{2}\vert\vert R \vert\vert^{2}}t^{-\frac{1}{2}\vert\vert R^{T} \vert \vert^{2}} M_{R}(U;q,t)M_{R}(V;q,t) \end{equation} where \(\vert \vert R \vert \vert^{2}= \sum_{i}R_{i}^{2}\) and \(\vert \vert R^{T}_{i} \vert \vert^{2} = \sum_{i} (R^{T}_{i})^{2} = \sum (2i-1)R_{i}\). \begin{figure}[htp] \centering \includegraphics[scale=0.80]{tqftpieces.pdf} \caption{The building blocks of the refined TQFT are the pant, cap, and annulus geometries, along with complex bundles of degree $(d_{1},d_{2})$ over each Riemann surface. \label{fig:tqftpieces}} \end{figure} Next we study the cap geometries (shown in Figure \ref{fig:tqftpieces}), which are given by two complex line bundles over the disc. Since the euler characteristic of the disc is equal to \(1\), the degrees of the line bundles must satisfy \(\textrm{deg}(\mathcal{L}_{1}) + \textrm{deg}(\mathcal{L}_{2}) = -1\). In practice, it suffices to determine the \((0,-1)\) amplitudes, since the rest can be obtained by gluing. It is helpful to notice that this geometry is equivalent to \(\mathbb{C}^{3}\) with a stack of branes inserted on one leg of the vertex. Thus the cap amplitude can be computed by the refined topological vertex amplitude, \(C_{R\cdot\cdot}\), with branes on the \(q\)-leg or from refined Chern-Simons \cite{AS2}. The result is, \begin{eqnarray} Z(C^{(0,-1)}) & = & \sum_{R}\frac{1}{g_{R}}\textrm{dim}_{q,t}(R)M_{R}(U;q,t) \end{eqnarray} where we have defined the \((q,t)\)-dimension of a representation \(R\) by, \begin{eqnarray} \textrm{dim}_{q,t}(R) & = & (q/t)^{\frac{1}{2}\vert R \vert}M_{R}(t^{\rho};q,t) \label{eqn:uinftyqtdim} \\ & = & q^{\frac{1}{4}\vert\vert R \vert\vert^{2}}t^{-\frac{1}{4}\vert\vert R^{T} \vert \vert^{2}} \prod_{\tableau{1}\in R}\Big(q^{\frac{a(\tableau{1})}{2}}t^{\frac{l(\tableau{1})+1}{2}}-q^{-\frac{a(\tableau{1})}{2}}t^{-\frac{l(\tableau{1})+1}{2}}\Big)^{-1} \nonumber \end{eqnarray} where \((\rho)_{i}=-i+1/2\) is the \(U(\infty)\) Weyl vector, while \begin{eqnarray} a(\tableau{1}) & = & R_{i}-j \label{eqn:armlength} \\ l(\tableau{1}) & = & R^{T}_{j}-i \nonumber \end{eqnarray} are the arm- and leg-lengths, respectively, of a box in the Young Tableau of \(R\). The \((q,t)\)-dimension can be understood as a \((q,t)\)-deformation of the dimension of the symmetric group representation specified by \(R\).\footnote{Our notation differs slightly from the notation used for the unrefined case in \cite{Aganagic:2004js}. In the limit \(t=q\), our \((q,t)\)-dimension is related to their \(d_{q}(R)\) by, \(\textrm{dim}_{q,q}(R) = s_{R}(q^{\rho}) = q^{\frac{1}{4}\kappa_{R}}d_{q}(R)\).} To obtain the cap with a different choice of line bundles we can simply glue on the \(A^{(1,-1)}\) or \(A^{(-1,1)}\) annuli. Finally, we must specify the three-punctured sphere amplitude (see Figure \ref{fig:tqftpieces}), which we refer to as the ``pant.'' Since the three-punctured sphere has the euler characteristic \(\chi=-1\), the degree of the line bundles must add to one in this case. To compute this amplitude it is helpful to recall some general properties that our TQFT must satisfy. Since it computes the refined topological A-model, the TQFT must be independent of complex structure moduli. Specifically this means that the amplitude for a Riemann surface should not depend on how it is formed by gluing simpler geometries. For this to be true, the pant amplitude should be symmetric in the three punctures, and thus should be diagonal in the Macdonald basis, \begin{equation} Z_{\textrm{ref top}}(P^{(0,1)}) = \sum_{R}P_{R}\;M_{R}(U_{1})M_{R}(U_{2})M_{R}(U_{3}) \end{equation} Now it is helpful to recognize that the pant, cap, and annulus are not all independent. We can form the annulus by capping off one of the punctures of the pant. By consistency and using the fact that \(Z(P^{(0,1)})\) is diagonal, we can solve for the pant amplitude, \begin{equation} Z_{\textrm{ref top}}(P^{(0,1)}) = \sum_{R}\frac{1}{g_{R}\textrm{dim}_{q,t}(R)}M_{R}(U_{1})M_{R}(U_{2})M_{R}(U_{3}) \end{equation} So far we have described the structure of the A-model on these geometries as a TQFT, but it is important to remember that the theory is not purely topological since it depends on the K\"ahler moduli. For Calabi-Yaus of the form \(\mathcal{L}_{1}\oplus \mathcal{L}_{2} \to \Sigma_{g}\), there is only one K\"ahler modulus, \(k\), that measures the area of the Riemann surface. In fact, as is familiar from the topological vertex \cite{Aganagic:2003db}, the partition function depends on this modulus only by introducing a term, \(e^{-k\vert R \vert}\) in the sum over representations. Altogether, we have given the necessary data to solve the theory completely. As an application of these results, we can study geometries of the form \(\mathcal{L}_{1}\oplus \mathcal{L}_{2} \to \Sigma_{g}\) where \(\Sigma_{g}\) is a genus \(g\) Riemann surface. For this to be a Calabi-Yau manifold we must have \(\textrm{deg}(\mathcal{L}_{1}) = 2g-2+p\) and \(\textrm{deg}(\mathcal{L}_{2}) = -p\). Then the refined amplitude on this geometry is given by, \begin{equation} Z_{\textrm{ref top}}^{(g,p)}(q,t) = \sum_{R}\Bigg(\frac{\textrm{dim}_{q,t}(R)^{2}}{g_{R}}\Bigg)^{1-g}q^{\frac{(2g-2+p)}{2}\vert\vert R \vert \vert^{2}}t^{-\frac{(2g-2+p)}{2}\vert\vert R^{T} \vert \vert^{2}}Q^{\vert R \vert} \label{eqn:reftopgeneral} \end{equation} where we have defined the exponentiated K\"ahler modulus as \(Q=e^{-k}\). It can be checked that this has the expected symmetry, \begin{equation} Z_{\textrm{ref top}}^{(g,p)}(q,t) = Z_{\textrm{ref top}}^{(g,p)}(t^{-1},q^{-1}) \end{equation} which implies that the Gopakumar-Vafa invariants come in complete multiplets of $SU(2)_\ell$ (as was the case in the unrefined limit). However, the amplitude is not symmetric under the exchange \(q\leftrightarrow t\). This implies that the Gopakumar-Vafa invariants for these Calabi-Yaus do not come in full representations of \(SU(2)_{r}\), but only carry \(U(1)_{r}\subset SU(2)_r\) charge. The BPS states come from quantizing the moduli space of curves in $X$, together with the $U(1)$ bundle on them. The $SU(2)_r$ spin content comes from cohomologies of the moduli of the curve itself, while the $SU(2)_{\ell}$ comes from the bundle. In the present case, the curve is $\Sigma$ itself. Its moduli space is in general non-compact, as typically one of the two line bundles over $\Sigma$ has positive degree. Correspondingly, the Lefshetz $SU(2)_r$ action on the cohomologies of the moduli space does not have to result in complete multiplets -- there can be contributions that escape to infinity. (See Appendix \ref{sec:gv} for some sample computations of Gopakumar-Vafa invariants for these geometries.) In fact, the only case when the moduli space is compact is when $\Sigma = {\mathbb P}^1$, and both line bundles are ${\cal O}(-1)$. It is easy to see that in this case the amplitude does in fact have the \(q\leftrightarrow t\) symmetry as well. In addition, the amplitude is not symmetric under exchange of the two line bundles, which is equivalent to taking \(p \to 2-2g-p\). This tells us that one of the line bundles is distinguished from the other in the refined setting. In fact, this arises because the index in equation \ref{eqn:refindex} includes an \(R\)-symmetry twist that rotates a specific bundle in the noncompact Calabi-Yau (for more details see \cite{AS2}). Equivalently, as will be explained in section \ref{sec:factor}, these refined topological string amplitudes can be obtained by taking the large N limit of \(D4\) branes wrapping one of the bundles. In the unrefined case, the large N limit does not retain information about which bundle the \(D4\) branes wrapped, but in the refined case this choice has an effect on the closed string amplitude. This is related to the above observation since in both the \(D4\) construction and in the closed string construction we must choose an R-symmetry rotation to preserve supersymmetry. However, these symmetries are not completely lost since the amplitude is symmetric under simultaneously exchanging \(q\leftrightarrow t\) and exchanging the bundles, \begin{equation} Z_{\textrm{ref top}}^{(g,p)}(q,t) = Z_{\textrm{ref top}}^{(g,2-2g-p)}(t,q) \end{equation} As we will explain in section \ref{sec:u1engineering}, \(p\) specifies the five-dimensional Chern-Simons coupling of the geometrically engineered gauge theory. In \cite{Awata:2008ed}, it was similarly observed that for geometries that engineer five-dimensional \(SU(N)\) gauge theories, the refined topological string is only symmetric under the simultaneous exchange of \(q\leftrightarrow t\) and inverting the Chern-Simons level, \(k\to -k\). So far, we have computed all the non-trivial contributions to the refined topological string. However, we should also include by hand the additional pieces that appear at genus zero and one. In the unrefined case, these arise from constant maps. In the refined case, for geometries that engineer five-dimensional gauge theories, these contributions arise from the classical prepotential and the one-loop determinant of the instanton partition function. These degree zero pieces take the form, \begin{equation} Z_{0}(q,t) = \Big(M(q,t)M(t,q)\Big)^{\chi/4}\exp\Bigg(\frac{1}{\epsilon_{1}\epsilon_{2}} \frac{a k^{3}}{6}+\frac{\epsilon_{2}}{\epsilon_{1}} b \frac{k}{24}\Bigg) \label{eqn:degzero} \end{equation} where \(M(q,t)\) is the refined MacMahon function, \begin{equation} M(q,t) = \prod_{i,j=1}^{\infty}\Big(1-t^{i}q^{j-1}\Big) \end{equation} and \(\chi\) is the euler characteristic of the Calabi-Yau, while \(a\) is related to the triple intersection of the K\"ahler class and \(b\) is the second Chern class of the Calabi-Yau. These numbers are ambiguous because of the non-compactness of our geometries, but it was argued in \cite{Aganagic:2004js} that for the connection with black holes the natural values are, \begin{equation} \chi = 2-2g, \qquad \qquad a = - \frac{1}{p(p+2g-2)}, \qquad \qquad b = \frac{p+2g-2}{p} \end{equation} Note, that we have split the MacMahon function into two pieces, related by interchanging \(q\) and \(t\). This split naturally appears in section \ref{sec:factor}, when making the connection with the refined black hole partition function. A similar splitting was recently observed for the motivic Donaldson-Thomas invariants of the conifold in \cite{Morrison:2011rz}. In section \ref{sec:u1engineering}, we will give further evidence that this refined amplitude is the correct one by comparing it with the equivariant instanton partition function of the geometrically engineered five-dimensional field theory. Before doing so, however, it will be helpful to discuss one final aspect of the TQFT that arises when D-branes are included in the fiber of the complex line bundles. \subsection{Branes in the Fiber \label{sec:fiberbranes}} So far we have solved for the refined string on bundles over closed Riemann surfaces and Riemann surfaces with boundaries. These boundaries naturally end on branes wrapping an \(S^{1}\) in the base and two dimensions in the fiber. However, for understanding the refined OSV conjecture, it will be helpful to also consider introducing branes in the fiber. For unrefined topological strings, this was studied in \cite{Aganagic:2004js}, and our analysis will follow a similar approach. We consider a lagrangian brane at a point, \(z\), in the base Riemann surface, \(\Sigma_{g}\). Since the brane is local in the base, we only need to study a neighborhood of \(z\). Thus it is natural to introduce branes in the base that chop up the geometry into a disc, \(D\), containing \(z\), and its complement, \(\Sigma \setminus D\). The full amplitude is given by, \begin{equation} Z = \sum_{R, Q}Z_{R}(\Sigma \setminus D)Z_{RQ}(D)M_{Q}(V;q,t) \end{equation} where \(V\) is the holonomy around the branes in the fiber. The amplitude on the complement, \(\Sigma\setminus D\), can be solved by gluing using the amplitudes in the previous section, but we still need to solve for the disc amplitude with two sets of branes. This can be accomplished by noticing that \(D\) has the topology of \(\mathbb{C}^{3}\) with the base and fiber branes on two legs of the vertex. Thus, the full cap amplitude, \begin{equation} Z(D) = \sum_{R,Q} Z_{RQ}(D)M_{R}(U)M_{Q}(V) \end{equation} is simply computed by the refined topological vertex amplitude with two stacks of branes on different legs, as shown in Figure \ref{fig:smat}. \begin{figure}[htp] \centering \includegraphics[scale=1.00]{Smat.pdf} \caption{The full cap amplitude with branes in both the fiber and the base. \label{fig:smat}} \end{figure} Alternatively, as will be explained in \cite{AS2}, this amplitude can be solved by following the refined Chern-Simons theory through a geometric transition. From this perspective, \(Z_{RQ}\) is computed by the large N limit of the refined Chern-Simons S-matrix, \begin{eqnarray} W_{RQ} & = & \lim_{N\to\infty}t^{-\frac{N(\vert R\vert + \vert Q \vert)}{2}}(q/t)^{\frac{\vert R \vert + \vert Q \vert}{4}}\frac{S_{RQ}(q,t;N)}{S_{00}(q,t;N)} \\ & = & (q/t)^{\frac{\vert R \vert + \vert Q \vert}{2}}M_{R}(t^{\rho})M_{Q}(t^{\rho}q^{R}) \end{eqnarray} where we have used the symmetrized definition of the infinite-variable Macdonald polynomials in Appendix \ref{sec:macs}. By including the appropriate metric factors, we obtain, \begin{equation} Z_{RQ} = \frac{1}{g_{R}g_{Q}}W_{RQ} \end{equation} Note that if we set the representation of the fiber brane to be trivial, \(Q=0\), then this geometry is the same as the cap (C) that we studied above. This is consistent with the fact that \(W_{R0} = \textrm{dim}_{q,t}(R)\). As an example, take the geometry, \(\mathcal{O}(2g-2+p)\oplus \mathcal{O}(-p) \to \Sigma_{g}\) with branes in the fiber over \(h\) points. Then the refined amplitude is given by, \begin{eqnarray} Z_{\textrm{ref top}}^{(g,p,h)}(q,t) = \sum_{R, R_{1}, \cdots, R_{h}}& & \frac{g_{R}^{g-1}}{W_{R0}^{2g-2+h}}\frac{W_{RR_{1}}\cdots W_{RR_{h}}}{g_{R_{1}}\cdots g_{R_{h}}}q^{\frac{(2g-2+p)}{2}\vert\vert R \vert \vert^{2}}t^{-\frac{(2g-2+p)}{2}\vert\vert R^{T} \vert \vert^{2}}Q^{\vert R \vert} \nonumber \\ & & \cdot M_{R_{1}}(V_{1})\cdots M_{R_{h}}(V_{h}) \end{eqnarray} It is also useful to understand how an anti-brane can be introduced that wraps the fiber. Recall that in the unrefined topological string, converting a brane into an anti-brane corresponds to taking, \begin{equation} s_{R}(U) \to (-1)^{\vert R \vert}s_{R^{T}}(U) \end{equation} where \(s_{R}(U)\) is the Schur function. The analogue of this reversal in the refined setting corresponds to taking, \begin{equation} M_{R}(U;q,t) \to \iota M_{R}(U;q,t) \end{equation} where \(\iota\) is defined by how it acts on power sums, \(p_{n}(x)\), \begin{equation} \iota(p_{n}) = -p_{n} \end{equation} We refer the reader to \cite{AS2, ASVertex} for more details on this construction. This implies that the disc amplitude with an anti-brane in the fiber is given by, \begin{equation} \widetilde{Z}(D) = \sum_{R,Q}\frac{1}{g_{R}g_{Q}}W_{RQ}M_{R}(U)\iota M_{Q}(V) \end{equation} If we want to rewrite this in the \(M_{Q}(V)\) basis, this can be done by using a generalized Cauchy identity (see Appendex \ref{sec:macs}), \begin{eqnarray} \widetilde{Z}(D) & = & \sum_{R,Q}\frac{1}{g_{R}g_{Q}}(q/t)^{\frac{\vert R \vert + \vert Q \vert}{2}}M_{R}(t^{\rho})M_{Q}(t^{\rho}q^{R})M_{R}(U)\iota M_{Q}(V) \\ & = & \sum_{R}\frac{1}{g_{R}}M_{R}(t^{\rho})M_{R}(U)\sum_{Q}\frac{1}{g_{Q}}(q/t)^{\frac{\vert R \vert + \vert Q \vert}{2}}M_{Q}(t^{\rho}q^{R})\iota M_{Q}(V) \\ & = & \sum_{R}\frac{1}{g_{R}}M_{R}(t^{\rho})M_{R}(U)\sum_{Q}\frac{1}{g_{Q}}(q/t)^{\frac{\vert R \vert + \vert Q \vert}{2}} \iota M_{Q}(t^{\rho}q^{R}) M_{Q}(V) \\ & = & \sum_{R,Q}\frac{1}{g_{R}g_{Q}}\widetilde{W}_{RQ}M_{R}(U)M_{Q}(V) \end{eqnarray} where we have defined \(\widetilde{W}_{RQ}\) by, \begin{equation} \widetilde{W}_{RQ} = (q/t)^{\frac{\vert R \vert + \vert Q \vert}{2}}M_{R}(t^{\rho})\iota M_{Q}(t^{\rho}q^{R}) \end{equation} This amplitude will be particularly important for studying the genus \(g=0\) OSV conjecture in section \ref{sec:genuszero}. It is important to note that the fiber brane has a modulus, \(k_{f}\). If we take \(u\) to be a coordinate for one of the fibers, then the fiber brane sits at \(\vert u \vert^{2}=const\). As is standard, this real modulus combines with the holonomy to form the complexified K\"ahler parameter, \(k_{f}\). Including this modulus simply modifies the partition function as, \begin{equation} M_{R}(U) \to e^{-k_{f}\vert R \vert}M_{R}(U) \end{equation} This modulus will appear in section \ref{sec:factor} when we discuss the ``ghost branes'' that appear in tests of the refined OSV conjecture. \subsection{Refined Topological Strings on $\mathcal{L}_{1}\oplus\mathcal{L}_{2}\to \Sigma_{g}$ and $5d$ $U(1)$ Gauge Theories\label{sec:u1engineering}} Now that we have defined a TQFT that computes refined topological string amplitudes, we would like to verify our proposal. A simple check is that the Gopakumar-Vafa invariants are integers. We have verified this in general, and we present a few examples in Appendix \ref{sec:gv}. We can perform a much stronger check of our proposal by using geometric engineering. We consider M-theory on the Calabi-Yau, \(X=\mathcal{O}(p)\oplus\mathcal{O}(2g-2-p)\to \Sigma_{g}\). This is known to engineer five dimensional \(U(1)\) gauge theory with \(g\) hypermultiplets in the adjoint representation, and with a level \(k_{CS}=1-g-p\) five-dimensional Chern-Simons term turned on \cite{Katz:1996ht, Chuang:2010ii, Chuang:2012dv}.\footnote{The motivic Donaldson-Thomas invariants of these geometries were also studied recently in \cite{Chuang:2010ii, Chuang:2012dv}. In general, the motivic invariants of a Calabi-Yau, \(X\), will differ from the refined invariants that we compute in this paper. Motivic invariants depend on the motive of X and thus are sensitive to its complex structure. In contrast, our refined invariants are computed by a supersymmetric index which makes them invariant under complex structure deformations. These differences are reflected in the connection with geometric engineering. In \cite{Chuang:2010ii, Chuang:2012dv}, the motivic invariants for these geometries were related to the instanton partition function with the adjoint mass equal to \(m=(\epsilon_{1}-\epsilon_{2})/2 \leftrightarrow \widetilde{y}=\sqrt{q/t}\). Our refined invariants are identified with the different parameter choice, \(m=0 \leftrightarrow \widetilde{y}=1\). We thank Emanuel Diaconescu for helpful discussions on this point.} Now we consider the K-theoretic equivariant instanton partition function for these theories.\footnote{Ordinarily, such counting would not be sensible because \(U(1)\) instantons are singular and because the adjoint representation of \(U(1)\) is trivial. However, this instanton counting is performed by turning on background noncommutativity which both resolves \(U(1)\) instantons and causes fields in the \(U(1)\) adjoint representation to transform nontrivially.} The original index in \cite{Nekrasov:2002qd} that computes the K-theoretic instanton partition function is exactly the same index that we have used to compute the refined A-model in equation \ref{eqn:refindex}, so the two partition functions must agree. As explained in \cite{Chuang:2012dv}, the instanton partition function for this five-dimensional field theory is given by, \begin{eqnarray} Z_{U(1)}^{g,k_{CS}}(q,t,\widetilde{Q}) & = & \sum_{\mu}\prod_{\tableau{1}\in \mu}\Big(q^{-l(\tableau{1})-1/2}t^{a(\tableau{1})+1/2}\Big)^{k_{CS}}\Big(1-q^{-l(\tableau{1})}t^{-a(\tableau{1})-1}\Big)^{g-1} \nonumber \\ & & \cdot\Big(1-q^{l(\tableau{1})+1}t^{a(\tableau{1})}\Big)^{g-1}(t/q)^{\frac{(g-1)\vert \mu \vert}{2}} \widetilde{Q}^{\vert \mu\vert} \end{eqnarray} where \(q\) and \(t^{-1}\) are the equivariant parameters rotating the \(z_{1}\) and \(z_{2}\) planes respectively. The sum is over all Young Tableaux, \(\mu\), and the arm and leg length of a box in such a tableau (defined in equation \ref{eqn:armlength}) are denoted by \(a(\tableau{1})\) and \(l(\tableau{1})\), respectively. By using the definitions of the metric and $(q,t)$-dimension in equations \ref{eqn:uinftymetric} and \ref{eqn:uinftyqtdim}, we can rewrite the equivariant instanton partition function as, \begin{eqnarray} Z_{U(1)}^{g,k_{CS}}(q,t,\widetilde{Q}) & = & \sum_{\mu}\Bigg(\frac{\textrm{dim}_{q,t}(\mu)^{2}}{g(\mu)}\Bigg)^{1-g} \Big(q^{-\frac{1}{2}\vert\vert \mu \vert \vert^{2}}t^{\frac{1}{2}\vert\vert \mu^{T} \vert \vert^{2}}\Big)^{k_{CS}+1-g}(-1)^{\vert R\vert}\widetilde{Q}^{\vert \mu \vert} \end{eqnarray} But now it is clear that this agrees with the refined topological string partition function of equation \ref{eqn:reftopgeneral}, \begin{equation} Z_{U(1)}^{g,k_{CS}}(q,t,\widetilde{Q}) = Z_{\textrm{ref top}}^{g,p}(q,t,Q) \end{equation} upon making the change of variables, \begin{eqnarray} \widetilde{Q} & = & (-1)^{g-1}Q \nonumber \\ k_{CS} & = & 1-g-p \nonumber \end{eqnarray} This verifies in general our proposed refinement of the Bryan-Pandharipande TQFT for arbitrary line bundles over a Riemann surface. \section{Refined Black Hole Entropy \label{sec:branes}} In this section we study BPS bound states of $N$ D4 branes wrapping a four-cycle inside a Calabi-Yau, and carrying D2 and D0 brane charge. We start by explaining that the refined counting of D4/D2/D0-brane BPS bound states computes the \(\chi_{y}\)-genus of the moduli space of instantons on the four-cycle wrapped by the D4 branes. We then specialize to the case of interest in this paper -- IIA string theory compactified to four-dimensions on the class of Calabi-Yau manifolds, \(X\), that consist of two complex line bundles over a Riemann surface, and show how to compute the $\chi_y$ genus in the examples that arise there. Finally, as a check of our results in this section, we compare our answers in the case when the four-cycle is ${\cal O}(-1)\rightarrow {\mathbb P}^1$ against a direct computation of the cohomologies of the moduli space of instantons, by Yoshioka and Nakajima in \cite{Yoshioka:1996pd, Nakajima:2003uh}, and find a perfect agreement. In the unrefined case, the black hole partition function is the index \begin{equation} Z_{BH} = \textrm{Tr}_{\mathcal{H}_{BPS}}(-1)^{F}e^{-\phi_{2}Q_{2}}e^{-\phi_{0}Q_{0}} \label{eqn:urtrace} \end{equation} where \(Q_{0}\) and \(Q_{2}\) are the D0 and D2 charges, while \(\phi_{0}\) and \(\phi_{2}\) are the respective chemical potentials. Since we are working in the large volume limit, we can identify D0/D2/D4 bound states with nontrivial \(U(N)\) bundles, \(V\), over \(C_{4}\). The D-brane charges and Chern classes of this bundle are related by, \begin{equation} Q_{2} = c_{1}(V), \qquad \qquad Q_{0} = ch_{2}(V) \end{equation} Therefore, calculating degeneracies will reduce to field theoretic computations on the D4-brane worldvolume. Since the D4 brane wraps \(\mathbb{R}_{t} \times C_{4}\), we can associate to \(C_{4}\) a Hilbert space\(\mathcal{H}\), which is graded by D2/D0-brane charge, angular momentum, \(J_{3}\), and R-charge, \(R\). Now we would like to compute the BPS degeneracies as a trace over this entire Hilbert space. This can be done easily by using the Witten index, since non-BPS contributions will cancel out. Therefore we must simply compute the D4-brane path integral on \(S^{1} \times C_{4}\), \begin{equation} Z_{BH} = \textrm{Tr}_{\mathcal{H}}(-1)^{F}e^{-\phi_{2}c_{1}}e^{-\phi_{0}ch_{2}} \end{equation} Since the D-branes are wrapping a curved geometry, the gauge theory is topologically twisted along \(C_{4}\)\cite{Bershadsky:1995qy}. In our case, \(\mathcal{H}_{BPS}\) is equal to the cohomology of instanton moduli space for the corresponding topological sector. Therefore, computing the Witten index reduces to computing the euler characteristic, \(\chi(\mathcal{M})\), for the moduli space of instantons on \(C_{4}\). We have presented this computation entirely from a five-dimensional perspective because this approach will easily generalize to the refined setting. However, in the unrefined case we could also reduce on the \(S^{1}\) and study the four-dimensional gauge theory. This leads to four-dimensional topologically twisted \(\mathcal{N}=4\) Yang-Mills \cite{Vafa:1994tf} on \(C_{4}\) with the observables, \begin{equation} S = \frac{\phi_{0}}{8\pi^{2}}\int {\rm Tr\,} F\wedge F + \frac{\phi_{2}}{2\pi}\int {\rm Tr\,} F \wedge \omega_{\Sigma} \end{equation} inserted into the action. Here, \(\omega_{\Sigma}\) is the K\"ahler class of the Riemann surface, \(\Sigma_{g}\). Since this is the Vafa-Witten \cite{Vafa:1994tf} twist of \(\mathcal{N}=4\), the four-dimensional perspective explains why we are computing the euler characteristic of instanton moduli space. Now that we have discussed the unrefined case, we would like to count BPS states while keeping information about angular momentum and R-charge. As explained in section \ref{sec:osv}, our goal is to compute the protected spin character of D2 and D0-branes bound to the D4-branes, \begin{equation} Z_{BH} = \textrm{Tr}_{\mathcal{H}_{BPS}}(-1)^{2J_{3}}y^{J_{3}-R}e^{-\phi_{2}Q_{2}}e^{-\phi_{0}Q_{0}} \label{eqn:rtrace} \end{equation} where we have used the variable, \(y\equiv e^{-2\gamma}\). Since this is an index, it only receives contributions from BPS states. This means we can extend the trace to be over the full D4-brane Hilbert space ${\cal H}$, \begin{equation} Z_{BH} = \textrm{Tr}_{\mathcal{H}}(-1)^{2J_{3}}y^{J_{3}-R}e^{-\phi_{2}Q_{2}}e^{-\phi_{0}Q_{0}} \label{eqn:rtrace}. \end{equation} This also means that $Z_{BH}$ can be computed by the five-dimensional path integral and will be invariant under small deformations.\footnote{One should contrast this with the most general trace, $$ Z_{BH} = \textrm{Tr}_{\mathcal{H}_{BPS}}(-x)^{J_{3}+R}(-y)^{J_{3}-R}e^{\phi_{2}Q_{2}-\phi_{0}Q_{0}} \label{eqn:double} $$ where \(J_{3}\) is the generator of the \(Spin(3)\) rotation group in the (3+1)-dimensional spacetime, and \(R\) is the \(U(1)\) R-charge of the four-dimensional BPS states. Unfortunately, this trace cannot be extended to the full Hilbert space, since non-BPS states will contribute nontrivially. This means that the doubly-refined trace in equation \ref{eqn:double} cannot be computed by a five-dimensional path-integral, and is therefore analogous to the five-dimensional Khovanov-Rhozansky construction of \cite{Witten:2011zz}.} To understand precisely what the protected spin character computes, it helps to remember that the Hilbert space \(\mathcal{H}_{BPS}\) can be identified with the cohomology of the moduli space \(\mathcal{M}\) of instantons on \(C_{4}\). Once we fix the topological charges \(c_{1}\) and \(ch_{2}\), the most general geometric quantity that can be computed from the cohomology of \(\mathcal{M}\) is the Hodge polynomial, \begin{equation} e(\mathcal{M};x,y) = \sum_{p,q}(-1)^{p+q}x^{p}y^{q}\textrm{dim} H^{p,q}(\mathcal{M}) \end{equation} As explained in \cite{Diaconescu:2007bf}, the degrees, \((p,q)\) are related to the R-charge and spin by, \begin{equation} J_{3} = \frac{p+q}{2}, \qquad \qquad R = \frac{p-q}{2} \label{eqn:hodgespin} \end{equation} Therefore, the refined black hole partition function in equation \ref{eqn:rtrace} computes the generating function for the \(\chi_{y}\) genus of instanton moduli space, \begin{eqnarray} Z_{BH} & = & \sum_{c_{1},ch_{2}}e^{-\phi_{0}ch_{2}-\phi_{2}c_{1}}\sum_{p,q}(-1)^{p+q}y^{q}h^{p,q}(\mathcal{M}_{c_{1},ch_{2}}) \\ & = & \sum_{c_{1},ch_{2}}e^{-\phi_{0}ch_{2}-\phi_{2}c_{1}}\chi_{y}(\mathcal{M}_{c_{1},ch_{2}}) \end{eqnarray} One more aspect of the protected spin character that we will need is its transformation properties under $S$-duality. In the unrefined case, the transformation properties are well known \cite{Vafa:1994tf}. The partition function \ref{eqn:urtrace} transforms like a theta function, with modular parameter $\phi_0$. We show in appendix \ref{sec:sduality} that in the refined case $S$-duality corresponds to replacing \begin{equation}\label{Sduality} \phi_{0} \to \frac{4\pi^{2}}{\phi_{0}}, \qquad \phi_{2} \to 2\pi i \frac{\phi_{2}}{\phi_{0}}, \qquad \gamma \to 2\pi i\frac{\gamma}{\phi_{0}} \end{equation} where the variable \(y\) in the \(\chi_{y}\) genus is related to \(\gamma\) by \(y = e^{-2\gamma}\). In the rest of this section, we will show that, in the simple example of the family of Calabi-Yaus we have been studying, the $\chi_y$ genus of the instanton moduli space is computable explicitly in terms of a topological theory on the base Riemann surface $\Sigma$. \subsection{D4 branes on ${\mathcal L}_1\oplus {\mathcal L}_2\rightarrow \Sigma$ \label{sec:d4x}} In our local Calabi-Yau manifold, $${\mathcal L}_1\oplus {\mathcal L}_2\rightarrow \Sigma_g$$ consider $N$ D4 branes wrapping the zero section of ${\mathcal L}_2$. The world-volume of the brane is $${\cal D} = ({\mathcal L}_1\to \Sigma_{g})$$ As before, we take ${\cal L}_1$ to have the first Chern class $-p$, so ${\mathcal L}_1$ is an $\mathcal{O}(-p)$ bundle over $\Sigma$. In the unrefined case studied in\cite{Aganagic:2004js}, the partition function of the Vafa-Witten twisted \(\mathcal{N}=4\) $U(N)$ Yang-Mills on ${\cal D}$ was shown to be computed by $q$-deformed two-dimensional bosonic Yang-Mills on \(\Sigma_{g}\). Roughly speaking, one can use localization on the fiber over the Riemann surface to reduce the four-dimensional theory down to a theory on $ \Sigma$. The basic observation is that one can use localization along the fiber of \(\mathcal{O}(-p) \to \Sigma_{g} \) to reduce the four dimensional theory to a two dimensional theory on the fixed point set. This reduces the observables, \begin{equation} S =\frac{\phi_0}{8\pi^{2}} \int {\rm Tr\,} F\wedge F + \frac{\phi_2}{2\pi} \int {\rm Tr\,} F \wedge \omega_{\Sigma} \end{equation} to \begin{equation} S = \frac{\phi_0}{4\pi^{2}} \int_{\Sigma_g}{\rm Tr\,} \Phi \; F + \frac{\phi_2}{2\pi} \int_{\Sigma_g} {\rm Tr\,} \Phi\; \omega_{\Sigma} - p \frac{\phi_0}{8\pi^{2}} \int_{\Sigma_g} \; {\rm Tr\,} \Phi^2 , \end{equation} which is the action of the bosonic two dimensional Yang-Mills. Here, $\Phi$ is the holonomy of the circle at infinity of the $\mathbb{C}$ fiber -- the action becomes a boundary term. The last term reflects the topology of the fibration over $\Sigma_g$. The way it arises from four dimensions was explained in \cite{Aganagic:2004js}. This is not quite the end of the story, as one has to be careful about the measure of the path integral. The fact that $\Phi$ comes from the holonomy around the $S^1$ turns it into a periodic variable -- this is why the theory is $q$-deformed 2d Yang-Mills, instead of ordinary 2d Yang Mills. In the limit where the \(D2\) brane chemical potential $\phi_2$ is turned off, the \(q\)-deformed Yang-Mills on the above geometry reduces to an analytic continuation of ordinary Chern-Simons theory on a degree \(p\) \(S^{1}\) bundle over \(\Sigma_{g}\). It is important to note that the ${\cal N}=4$ YM and Chern-Simons couplings are the same. In this paper, we would like to solve for the corresponding refined amplitudes. Since all of the arguments in the derivation so far were topological, the only thing that can change in a non-trivial way is the measure of the two-dimensional path integral. While deriving the measure is straightforward in the unrefined case, it is more challenging in the refined theory. Instead, we will give pursue a different path. We will find another derivation of the fact that the Vafa-Witten partition function in this background is computed by $q$-deformed 2d Yang-Mills, in which understanding the deformation we need will be easy. It will turn out that the theory we get is related to the refined Chern-Simons theory of \cite{Aganagic:2011sg, AS2}. \subsection{From 4d ${\cal N}=4$ Yang-Mills to 2d $(q,t)$-deformed Yang-Mills \label{sec:n4qt}} The idea of the derivation is to look at the same D4 brane background in a slightly different way -- by pairing the coordinates differently to get a Calabi-Yau four-fold instead. Doing so will make it manifest that, in the unrefined case, the theory we get is the same as Chern-Simons theory at $\phi_2=0$, or more generally, $2d$ $q$-deformed Yang Mills. To begin with, we will consider the case \(p=0\) case, so the Calabi-Yau manifold is simply $$ (\mathcal{O}(0) \oplus \mathcal{O}(2g-2) \to \Sigma_g ) = {\mathbb C} \times T^*\Sigma_g $$ and the D4 branes wrap the divisor $${\cal D} = ({\cal O}(0)\to \Sigma_{g}) = {\mathbb C} \times \Sigma_g.$$ Thus, all together, the Vafa-Witten theory we are interested in, as we explained above, arises from studying the partition function of the $N$ D4-branes wrapping, \begin{equation} \mathbb{C} \times \Sigma_{g} \times S_t^{1} \label{Ob} \end{equation} in IIA string theory on, \begin{equation} \mathbb{C} \times T^{*}\Sigma_{g}\times \mathbb{R}^{3} \times S_t^{1} .\label{eqn:qspacetime} \end{equation} As we go around the temporal circle, \(S_t^{1}\), we compute the index, \begin{equation} Z = \textrm{Tr}(-1)^{F}e^{- \phi_0 Q_{0}} \end{equation} where \(Q_{0}\) is the \(D0\)-brane charge bound to the \(D4\)-branes. We have temporarily set $\phi_2$ to zero. Our goal is to explain why this construction leads to the partition function of analytically continued Chern-Simons on \(S^{1} \times \Sigma_{g}\). Recall moreover that after a modular transformation, the partition function becomes manifestly equal to the partition function of Chern-Simons theory, with $q= e^{g_s} =e^{\frac{4\pi^2}{\phi_0}}$. In \cite{Witten:2011zz}, analytically continued $SU(N)$ Chern-Simons theory on $S^1\times \Sigma_g$ was obtained from a string theory construction involving a stack of $N$ D4 branes wrapping \begin{equation} \mathbb{C} \times \Sigma_{g}\times S_t^{1} \end{equation} in IIA string theory on the Calabi-Yau fourfold $ T^{*}(\mathbb{C}\times \Sigma_{g}) $, or more precisely, on \begin{equation} T^{*}(\mathbb{C}\times \Sigma_{g}) \times \mathbb{R} \times S_t^{1} = \mathbb{C} \times T^{*} \Sigma_{g} \times\mathbb{R}^{3} \times S_t^{1} \label{Wb}. \end{equation} At infinity of the D4 brane, we impose the D6 brane boundary conditions along \begin{equation} T^{*}(S^{1} \times \Sigma_{g}) \times \{0\} \times S^{1}_{t} \end{equation} We view the $\mathbb{C} \times \Sigma_{g}$ as the base of the cotangent space, and the D6 brane wraps the $S^1$ at the boundary of $\mathbb{C}$. The theory on the D4 branes (forgetting the temporal circle) on a Lagrangian cycle in a Calabi-Yau fourfold, is the Langlands twist of ${\cal N}=4$ $U(N)$ Yang-Mills on $ \mathbb{C} \times \Sigma_{g}$. Upon going around the first \(S^{1}\), we compute the index on the \(D4\) brane worldvolume, \begin{equation} Z = \textrm{Tr}(-1)^{F}e^{-\phi_0' Q_{0}} \label{eqn:wittenindex} \end{equation} where $\phi_0' = \frac{4\pi^2}{\phi_0}=g_s$. Using string dualities, \cite{Witten:2011zz} argued that the theory on the D4 branes is $U(N)$ Chern-Simons theory, with $q=e^{g_s}.$ It may be surprising at first, but these two constructions are effectively the same. In the case studied in \cite{Witten:2011zz} one has $\mathcal{N}=4$ theory with the Langlands twist, while we are a priori interested in the Vafa-Witten twist. While the two are not the same on a generic four manifold $V$, if we take $V=\mathbb{C} \times \Sigma_{g}$, the difference disappears. We can argue that this is the case by recalling that topologically twisting merely implements the twisted version of supersymmetry imposed by the string background. As is manifest from \ref{Ob} and \ref{Wb}, the string backgrounds end up being the same in our case. The other apparent difference is that the Witten construction involves \(D4\)-branes ending on a \(D6\)-brane at infinity. In contrast, the first construction naively only involves \(D4\)-branes. This discrepancy can be resolved by remembering that in the our setup, we still must impose boundary conditions at infinity on the noncompact \(\mathbb{C}\). If we were to choose the boundary conditions $S$-dual to those of the \(D6\)-brane boundary conditions for the \(S^{1}\) at infinity, then the two setups agree. The S-duality is here simply to account for the fact that with D6 brane boundary conditions it is $q=e^{g_s}$ that keeps track of the instanton charge, where $q$ is parameter in terms of which the one naturally writes the Chern-Simons amplitudes -- while with the S-dual boundary conditions instead, it is $e^{-1/g_s}$ that keeps track of the instanton charge in the gauge theory on the four-manifold. As preparation for understanding the refined theory (and to ultimately make contact with the definition of refined Chern-Simons in \cite{Aganagic:2011sg}), it is helpful to also consider the unrefined setup in a slightly different geometry. On very general grounds \cite{Witten:2011zz}, we expect the partition function of Langlands-twisted $SU(N)$ \(\mathcal{N}=4\) theory on a four-manifold \(V\), to be equal to the partition function of the $SU(N)$ Chern-Simons theory on the boundary \(\partial V\) of the four manifold. The choice of the bulk geometry, \(V\), only potentially affects the integration contour of the analytically continued Chern-Simons partition function. In the previous setup we studied \(V= \mathbb{C} \times \Sigma\), but instead we could choose \(V = \mathbb{R}_{+} \times S^{1} \times \Sigma\). More explicitly we could take IIA string theory on the geometry, \begin{equation} T^{*}({\mathbb C}^* \times \Sigma) \times \mathbb{R} \times S_{t}^{1} = \mathbb{C}^{*} \times T^{*} \Sigma \times\mathbb{R}^{3} \times S_t^{1} \label{W2} \end{equation} with \(D4\) branes wrapping \begin{equation} \mathbb{R}_{+}\times S^{1} \times \Sigma\times S^{1}_{t} \end{equation} and with D6 brane boundary condition along, \begin{equation} \{0\} \times T^{*}(S^{1}\times \Sigma) \times \{0\} \times S^{1}_{t}. \end{equation} This is the more familiar realization of Chern-Simons that appears in the study of topological strings and in \cite{Witten:2011zz}. Now we would like to understand the effect of refinement on these setups. Firstly, as we argued, the setups are effectively the same for the purposes of the index, so if we understand refinement in any one of these, we will have understood it in all the others as well. Moreover, note that already in the unrefined case, the first and the second (or third) setup, are related by $S$-duality. Thus, if, in the second and third constructions, we are computing the index \begin{equation} Z_{ref} = \textrm{Tr}(-1)^{F}\exp\Bigg(-\phi_0'Q_{0} - 2\gamma'(J_{3}-R)\Bigg), \end{equation} where \(J_{3}\) rotates the \(\mathbb{R}^{3}\) spacetime and the R-symmetry acts geometrically by rotating the fiber of \(T^{*}\Sigma_{g}\), the index in the first setup is related to this by S-duality, \eqref{Sduality} \begin{equation}\label{eq:smap} \phi_{0}' \;\;\to\;\; \phi_0 = \frac{4\pi^{2}}{\phi_{0}'}, \qquad \gamma' \;\;\to\;\; \gamma = 2\pi i \frac{\gamma'}{\phi_{0}'} \end{equation} and equals \begin{equation} Z_{ref} = \textrm{Tr}(-1)^{F}\exp\Bigg(-\phi_0Q_{0} - 2\gamma(J_{3}-R)\Bigg). \end{equation} This would in principle simply provide alternate setups to compute the index, but it would not save us the work of actually evaluating it. Fortunately, however, in the third setup, the index was already computed. The problem of evaluating the refined index in this context was solved in \cite{Aganagic:2011sg, Aganagic:2012au}, in terms of the refined Chern-Simons theory on \(S^{1} \times \Sigma_{g}\). Since all three different setups, with the identification of parameters as in \ref{eq:smap} give rise to the same partition function, we conclude that the partition function in the first setup is simply the refined Chern-Simons partition function! For $\gamma'=0$, refined Chern-Simons becomes the same as ordinary Chern-Simons theory, analytically continued away from the integer level. As shown in \cite{Aganagic:2004js}, this in turn is the same as the 2d $q$-deformed Yang-Mills theory on $\Sigma$, upon reduction on the $S^1$ factor. Thus, we have derived the result of \cite{Aganagic:2004js}, by different means. Moreover, we have explained how to generalize it to the refined case. We will explain below that there is a two dimensional theory theory related to refined Chern-Simons theory the same way the $q$-deformed Yang -Mills is related to the ordinary Chern-Simons theory; we will call this theory the $q,t$-deformed 2d Yang-Mills. The only thing that remains to do is identify $\phi_0$ and $\gamma$, with the parameters $q,t$ that appear in the refined Chern-Simons partition function. To do this, we need to back up slightly, and recall how the refined Chern-Simons theory was defined originally. The refined Chern-Simons partition function is defined as the index of M-theory in the background that arises by simply uplifting the third setup to M-theory. In this case, D6 branes lift to Taub-Nut space, and D4 branes lift to M5 branes. All together, we get M-theory on \begin{equation} T^*(S^1\times \Sigma)\times TN \times S^{1} ={\mathbb C}^* \times T^*\Sigma\times TN \times S^{1} \end{equation} with M5 branes on \begin{equation} S^1\times \Sigma\times {\mathbb C} \times S^{1} \end{equation} In addition, as we go around the $S^1$, the Taub-NUT is twisted by, \begin{eqnarray} (z_{1}\, , \, z_2 ) \qquad \to \qquad (q\, z_{1} , \,t^{-1}\,z_{2} ). \end{eqnarray} So, from this perspective, the index we are computing takes the form similar to that in equation \eqref{eqn:refindex}: \begin{equation} Z_{ref} = \textrm{Tr}(-1)^{F} q^{S_1 - R} t^{R-S_2} , \end{equation} where $S_1$ rotates the complex plane $z_1$ wrapped by the M5 brane and $S_2$ generates the rotation of the $z_2$ plane, transverse to M5 brane and $R$ is another R-symmetry, coming from rotations of the fiber of $T^*\Sigma$. Moreover, let $$ q=e^{-\epsilon_1} ,\qquad t=e^{-\epsilon_2}. $$ We will see that these will turn out to be exactly the $\epsilon_1$ and the $\epsilon_2$ parameters that arize in the refined topological string. If we reduce this to IIA, we get the third setup back with the definition of refined Chern-Simons we started the discussion with. Naively, following arguments similar to those in section 2, one would have expected that we simply have $e^{-\phi_0'} = \sqrt{qt}=e^{-\frac{\epsilon_1+\epsilon_2}{2}}$ while $e^{\gamma' }= \sqrt{q/t}=e^{\frac{\epsilon_1-\epsilon_2}{2}}$. However, this would have treated $q$ and $t$ symmetrically, while in refined Chern-Simons theory the symmetry is badly broken. The fact that it is broken is very natural from the M-theory perspective, as on the M5 brane wrapping the $z_1$ plane $S_1$ corresponds to angular momentum, while $S_2$ is an R-symmetry, rotating the space transverse to the brane. Correspondingly, had we considered the M5 brane wrapping the $z_2$ plane instead, $S_2$ would have been the momentum on the brane. To reconcile these two perspectives, one from M-theory with M5 branes and the other from IIA with D4 branes, we must understand how the choice of \(q/t\)-branes appears in the IIA geometry. From the geometry of Taub-NUT space, it can be seen that this corresponds to whether the \(D4\)-brane runs along the positive or negative half-line in \(\mathbb{R}^{3}\). More precisely, the choice of \(q/t\)-branes is translated into whether it fills, \begin{equation} \{0\} \times \{z>0\} \in \mathbb{R}^{3} \qquad \textrm{or} \qquad \{0\}\times \{z<0\} \in \mathbb{R}^{3} \end{equation} Now to see how this affects the identification of \(D0\)-brane charge, recall that D0-branes are magnetically dual to D6-branes. In the presence of both kind of branes, the electromagnetic fields carry one unit of angular momentum along the vector connecting their positions. But the D6-brane in our geometry is frozen at \(\{0\}\in \mathbb{R}^{3}\) and the D0-branes that bind to the \(D4\)-brane must sit at \(\{z>0\}\) or \(\{z<0\}\) depending on the type of refined brane. Therefore, we find that the D0-brane always carries either \(+\frac{1}{2}\) unit of angular momentum, \(J_{3}\), or \(-\frac{1}{2}\) unit of angular momentum, depending on the type of \(D4\)-branes used. When we compute the above trace, it is natural to only count the angular momentum that comes from other physics, and absorb this intrinsic angular momentum into the weighting of D0-brane charge. From equation this implies that each D0-brane is weighted by \begin{equation} \big(\sqrt{qt}\big)^{Q_{0}}\big(\sqrt{q/t}\big)^{Q_{0}} = q^{Q_{0}} \qquad \textrm{or} \qquad \big(\sqrt{qt}\big)^{Q_{0}}\big(\sqrt{q/t}\big)^{-Q_{0}} = t^{Q_{0}} \end{equation} in perfect agreement with the M-theory perspective. Putting $$ q=e^{-\epsilon_{1}}\qquad t=e^{-\epsilon_{2}}. $$ this implies we should identify (for the D4 brane ending from \(\{z>0\}\) on the D6 brane) $$ \phi_0'=\epsilon_1', \qquad \gamma'=\frac{\epsilon_1-\epsilon_2}{2}. $$ We can use this, and $S$-duality, to identify the parameters $\phi_0$ and $\gamma$ in terms of $q$ and $t$, as: $$ \phi_0=\frac {4 \pi ^2}{\epsilon_1}, \qquad \gamma= \pi i \frac{\epsilon_1-\epsilon_2}{\epsilon_1} $$ So far our discussion has focused on the \(p=0\) case. Once we consider nontrivial circle or line bundles over \(\Sigma\), the setups will differ since the Vafa-Witten and Langlands twists are not equivalent when \(p\neq 0\). As discussed in \cite{Aganagic:2012au}, the framing factors in refined Chern-Simons can be understood as arising from a topological term. We expect that such topological terms should be present, regardless of which setup we use. Finally, we should remember that our goal is to count both \(D0\)-brane charge and \(D2\)-brane charge. Therefore, we must reintroduce a term in the index, \(e^{-\phi_{2}Q_{2}}\). This term is unaffected by refinement and takes precisely the same form as before. \subsection{Refined Chern-Simons Theory and a $(q,t)$-deformed Yang-Mills \label{sec:path}} To summarize, the refined black hole partition function, \begin{equation} Z_{ref} = \textrm{Tr}(-1)^{F}\exp\Bigg(-\phi_0Q_{0} - 2\gamma(J_{3}-R)\Bigg). \end{equation} corresponding to $N$ D4 branes on the divisor $${\cal D} = ({\cal O}(-p)\to \Sigma_{g})$$ is computed by refined $U(N)$ Chern-Simons theory on an $S^1$ bundle over $\Sigma_g$ of first Chern-Class $-p$, where the $q=e^{-\epsilon_1},t=e^{-\epsilon_2}$ parameters of refined Chen-Simons are related to $\phi_0$ and $\gamma$ as \begin{equation} \epsilon_1 = \frac{4\pi^2}{\phi_{0}} , \qquad \epsilon_2 = {4\pi^2 \over \phi_0}(1-\frac{\gamma}{2\pi i}), \qquad \theta = \frac{2\pi\phi_{2}}{\phi_{0}} \label{eqn:cov} \end{equation} It will be useful for us to formulate the refined Chern-Simons theory as a two dimensional one, refining $q$-deformed 2d Yang-Mills, in particular since we still need to turn on $\phi_2$, the D2 brane chemical potential (we could have done this in the refined Chern-Simons theory as well using the natural contact structure, but we will not do that here). As we explained earlier, the topological terms in four dimensional action are unchanged by refinement; only the values of $\phi_0$, $\phi_2$ change. \begin{equation} S =\frac{\phi_0}{8 \pi^2} \int {\rm Tr \, }F\wedge F + \frac{\phi_2}{2\pi} \int {\rm Tr \, }F \wedge \omega_{\Sigma} \end{equation} Namely, localization on the ${\cal L}_1 = {\cal O}(-p)$ fiber, relates the 4d observables, to 2d ones \begin{equation} S = \frac{\phi_0}{4\pi^{2}} \int_{\Sigma_g}{\rm Tr \, } \Phi \; F + \frac{\phi_2}{2\pi} \int_{\Sigma_g}{\rm Tr \, } \Phi\; \omega_{\Sigma} - p\frac{\phi_{0}}{8\pi^{2}} \int_{\Sigma_g} {\rm Tr \, }\Phi^2\; \omega_{\Sigma}. \end{equation} Here, $\Phi$ is the holonomy of the four-dimensional gauge field around the circle at infinity of the ${\cal L}_1$ fiber. We can also think of it as the holonomy of the Chern-Simons gauge field around the $S^1$. The origin of the last term, from the Chern-Simons perspective was reviewed in \cite{Aganagic:2012ne, Aganagic:2012au}. Here $\omega_{\Sigma}$ is a volume form on $\Sigma_g$, normalized to unit volume. The presence of this form in the action means that the 2d YM is invariant under area preserving diffeomorphisms only. Thus, we still get the action of the ordinary 2d Yang-Mills, but the measure has to be deformed -- both because of the periodicity of $\Phi$, and now because also of the $q,t$ dependence of the index. The measure factor was in fact the only difference between the refined and ordinary Chern-Simons theory, as well. Let ${\cal D}_{q,t} A$ be the refined Chern-Simons measure. This induces a measure on the gauge fields in two dimensions, but also on the holonomy. We will explain what the measure is in some detail later on, for now, let us leave it schematic. All together, the path integral of the theory is $$ Z_{\textrm{ref BH}}=\int {\cal D}_{q,t} A\;{\cal D}_{q,t} \Phi\; \exp\Bigl( \frac{\phi_0}{4\pi^{2}} \int_{\Sigma_g}{\rm Tr \, } \Phi \; F + \frac{\phi_2}{2\pi} \int_{\Sigma_g}{\rm Tr \, } \Phi\; \omega_{\Sigma} - p\frac{\phi_{0}}{8\pi^{2}} \int_{\Sigma_g} {\rm Tr \, }\Phi^2\; \omega_{\Sigma}. \Bigr) $$ We will refer to this theory as \((q,t)\)-deformed Yang-Mills theory.\footnote{This \((q,t)\)-deformed Yang-Mills theory has also appeared as a limit of the TQFT that computes the four-dimensional \(\mathcal{N}=2\) superconformal index in \cite{Gadde:2011uv}. It is also worth noting that in the limit \(q\to 1, t \to 1\) this theory reduces to ordinary 2d Yang-Mills at zero coupling.} We will use this path integral to derive the answer for the partition function, in the next subsection. For now, let us simply state the answer: For $N$ D4 branes wrapping a degree \(-p\) complex line bundle fibered over a genus \(g\) Riemann surface \(\Sigma_{g}\), the resulting refined partition function for bound states with D2-D0 branes is given by, \begin{equation} Z_{\textrm{ref BH}}= \sum_{\mathcal{R}}\Bigg(\frac{(S_{\mathcal{R}0})^2}{G_{\mathcal{R}}}\Bigg)^{1-g} q^{\frac{p(\mathcal{R},\mathcal{R})}{2}}t^{p(\rho,\mathcal{R})}Q^{\sum_{i}\mathcal{R}_{i}} \label{eqn:qtym11} \end{equation} where the sum is over representations, \(\mathcal{R}\), of \(U(N)\), \(\rho\) is the Weyl vector \((\rho)_{i}=\frac{N+1}{2}-i\), and \(Q=e^{-\phi_{2}}\) is the \(D2\)-brane chemical potential. Here we have also used the definitions, \begin{eqnarray} S_{\mathcal{R}0} = S_{00}\textrm{dim}_{q,t}(\mathcal{R}) & = & \prod_{m=0}^{\beta-1}\prod_{1\leq i<j \leq N} \lbrack \mathcal{R}_{i} - \mathcal{R}_{j} + \beta(j-i)+m \rbrack_{q} \nonumber \\ G_{\mathcal{R}} & = & \prod_{m=0}^{\beta-1}\prod_{1 \leq i<j \leq N}\frac{ \lbrack \mathcal{R}_{i}-\mathcal{R}_{j}+\beta(j-i)+m\rbrack_{q} }{\lbrack \mathcal{R}_{i}-\mathcal{R}_{j}+\beta(j-i)-m\rbrack_{q}} \end{eqnarray} Note that \(G_{\mathcal{R}}\) is naturally the finite N version of the metric that appeared in equation \ref{eqn:infmetric}. \subsection{A Path Integral for $(q,t)$-deformed Yang-Mills \label{sec:path}} Let us now explain in more detain what the $q,t$-deformed 2d Yang -Mills is, and how to compute its partition function. The most straight-forward way to proceed is to note that, because the theory is essentially topological, we can simply formulate the theory on pieces of the Riemann surface, explain how to glue them together, and show that the answer is independent of the decomposition. This has essentially been done in \cite{Aganagic:2011sg, Aganagic:2012ne}, only from the 3d perspective of the refined Chern-Simons theory on $S^1$ fibration over the Riemann surface. To avoid simply repeating the derivation of \cite{Aganagic:2011sg, Aganagic:2012ne}, we will instead compute the path integral directly. We will begin by recalling some of the results of \cite{Aganagic:2004js}, where the \(U(N)\) \(q\)-deformed 2d Yang-Mills theory was studied. This section will not be entirely self contained, but will build on \cite{Aganagic:2004js}. In \cite{Aganagic:2004js}, it was shown that the path integral of the unrefined theory can be abelianized, so that we are left with \(U(1)^{N}\) gauge fields, \(A_{k}\), and \(N\) compact scalar fields, \(\phi_{k}\), which are the eigenvalues of $\Phi$. Starting from the original integral: $$ Z_{\textrm{ BH}}=\int {\cal D} A\;{\cal D} \Phi\; \exp\Bigl( \frac{1}{g_s} \int {\rm Tr \, } \Phi \; F + \frac{\theta}{g_s} \int{\rm Tr \, } \Phi\; \omega_{\Sigma} - \frac{p}{g_s} \int \; {\rm Tr \, } \Phi^2\; \omega_{\Sigma}. \Bigr) $$ the abelianized version becomes \begin{equation} Z_{\textrm{qYM}}= \frac{1}{N!}\int^{'}\prod_{i}\mathcal{D}\phi_{i}\mathcal{D}A_{i} \Big(\Delta(\phi)\Big)^{1-g}\exp\Bigg(\sum_{i} {1\over g_s} \int_{\Sigma}d^{2}\sigma \Big(\frac{p}{2}\;\phi_{i}^{2}-{\theta} \;\phi_{i}\Big) -\frac{1}{g_s} \int_{\Sigma} F_{i}\phi_{i}\Bigg) \label{eqn:path} \end{equation} where we have used the measure factor, \begin{equation} \Delta(\phi) = \prod_{i\neq j}\Big(e^{\frac{\phi_{i}-\phi_{j}}{2}}-e^{\frac{\phi_{j}-\phi_{i}}{2}} \Big) \end{equation} and where \(\int^{'}\) indicates that the path integral omits those values of \(\phi\) for which \(\Delta(\phi)=0\). As was argued in \cite{Aganagic:2004js}, this partition function is precisely equal to the black hole partition function of equation \ref{eqn:urtrace} with the identification, \begin{equation} \phi_{0} = \frac{4\pi^{2}}{g_s}, \qquad \qquad \phi_{2} = \frac{2\pi \theta}{g_s} \end{equation} Since this action is quadratic and very simple, the path integral of $q$-deformed Yang-Mills can be solved exactly. As we discussed, in the refined case, the only thing that changes are the chemical potentials and the measure factors. The chemical potentials become \begin{equation} \phi_{0} = \frac{4\pi^{2}}{\epsilon_1}, \qquad \qquad \phi_{2} = \frac{2\pi \theta}{\epsilon_1} \end{equation} and $\gamma$ enters through the measure of the path integral that depends on both $\epsilon_1$ and $\epsilon_2$: \begin{equation} \epsilon_2 = {4\pi^2 \over \phi_0}(1-\frac{\gamma}{2\pi}) . \end{equation} In the refined theory, the measure becomes \begin{equation} \Delta(\phi) \to \Delta_{q,t}(\phi) = \prod_{m=0}^{\beta-1}\prod_{j\neq k} \Big(q^{-{\frac{m}{2}}}e^{\frac{\phi_{i}-\phi_{j}}{2}}- q^{{\frac{m}{2}}}e^{\frac{\phi_{j}-\phi_{i}}{2}} \Big) \label{eqn:measure} \end{equation} where \(\beta=\epsilon_{2}/\epsilon_{1}\), and we have taken it to be a positive integer for computational convenience. The path integral is thus given by: \begin{equation} Z_{qt\textrm{YM}}(\Sigma) = \frac{1}{N!}\int^{'}\prod_{i}\mathcal{D}\phi_{i}\mathcal{D}A_{i} \Big(\Delta_{q,t}(\phi)\Big)^{1-g}\exp\Bigg(\sum_{i} \int_{\Sigma}d^{2}\sigma\Big(\frac{p}{2\epsilon_{1}}\phi_{i}^{2}-\frac{\theta}{\epsilon_{1}} \phi_{i}\Big)-\frac{1}{\epsilon_{1}} \int_{\Sigma} F_{i}\phi_{i}\Bigg) \end{equation} Since this path integral is abelian, we can evaluate it explicitly following the approach of \cite{Blau:1993tv, Aganagic:2004js}. We begin by evaluating the path integral over abelian gauge fields. It is helpful to first change integration variables from \(\mathcal{D}A_{i}\) to \(\mathcal{D}F_{i}\). However, we should only integrate over those two-forms \(F_{k}\) that are genuine bundles over \(\Sigma_{g}\), which means that we must impose, \begin{equation} \int_{\Sigma_{g}}F_{k} \in 2\pi \mathbb{Z} \end{equation} This can be accomplished by inserting a periodic delta function, \begin{equation} \sum_{\{n_{k}\}}\delta^{(N)}\Big(\int_{\Sigma_{g}}F_{k} - 2\pi n_{k} \Big) \end{equation} which can be rewritten using Poisson resummation, \begin{equation} \sum_{\{n_{k}\}}\delta^{(N)}\Big(\int_{\Sigma_{g}}F_{k} - 2\pi n_{k} \Big) = \sum_{\{n_{k}\}}\exp\Big(in_{k}\int_{\Sigma_{g}}F_{k}\Big) \end{equation} Therefore the path integral takes the form, \begin{eqnarray} Z_{qt\textrm{YM}}(\Sigma) = \frac{1}{N!}\int^{'} & & \prod_{i}\mathcal{D}\phi_{k}\mathcal{D}F_{k} \Big(\Delta_{q,t}(\phi)\Big)^{1-g} \\ & & \cdot\exp\Bigg(\sum_{i} \int_{\Sigma}d^{2}\sigma\Big(\frac{p}{2\epsilon_{1}}\phi_{k}^{2}-\frac{\theta}{\epsilon_{1}} \phi_{k}\Big)-\frac{1}{\epsilon_{1}} \int_{\Sigma} F_{k}(\phi_{k} - i\epsilon_{1}n_{k}) \Bigg) \nonumber \end{eqnarray} When performing the path integral over \(F_{k}\) we obtain a delta function from the last term, and the path integral over \(\phi\) only receives contributions from constant fields, \begin{equation} \phi_{k} = i\epsilon_{1}n_{k} \end{equation} Therefore, the path integral evaluates to, \begin{equation} Z_{qt\textrm{YM}}(\Sigma) = \frac{1}{N!}\sum_{\{n_{k}\}}'\Big(\Delta_{q,t}(i\epsilon_{1}n_{k})\Big)^{1-g}\exp\Big(-\frac{p\epsilon_{1}}{2}\sum_{k}n_{k}^{2}-i\theta \sum_{k} n_{k}\Big) \label{eqn:qtym1} \end{equation} where \(\sum'\) indicates that we should only sum over those field configurations where \(\Delta_{\epsilon_{1},\beta}(\phi)\) is nonzero. From equation \ref{eqn:measure}, this means that we must impose \(n_{i} \neq n_{j}\), but we also require, \begin{eqnarray} n_{i} & \neq & n_{j} \pm 1 \nonumber \\ n_{i} & \neq & n_{j} \pm 2 \label{eqn:constraints} \\ & \vdots & \nonumber \\ n_{i} & \neq & n_{j} \pm (\beta-1) \nonumber \end{eqnarray} It is also helpful to notice that both the refined measure and the action are invariant under Weyl reflections, so we can restrict the sum to the fundamental Weyl chamber so that \(n_{1}> n_{2} > \ldots > n_{N}\). We would like to use these observations to rewrite equation \ref{eqn:qtym1} as a sum over \(U(N)\) representations. Recall that a \(U(N)\) representation is specified by highest weights satisfying \(\mathcal{R}_{1} \geq \mathcal{R}_{2} \geq \ldots \geq \mathcal{R}_{N}\). Therefore, in order to satisfy the constraints of equation \ref{eqn:constraints}, we should shift each \(n_{k}\) by \(\beta k\). For convenience, we can also shift all \(n_{k}\) by a constant amount. This leads us to the identification, \begin{equation} n_{k} = \mathcal{R}_{k}+\beta \rho_{k} \end{equation} where \(\rho_{k} = \frac{N+1}{2}-k\). Then the partition function takes the form, \begin{equation} Z_{qt\textrm{YM}}(\Sigma) = \sum_{\mathcal{R}}\Big(\Delta_{q,t}(\mathcal{R})\Big)^{1-g} q^{\frac{p}{2}(\mathcal{R},\mathcal{R})}t^{p(\rho,\mathcal{R})}Q^{\vert \mathcal{R} \vert} \end{equation} where we have defined \(q=e^{-\epsilon_{1}}\), \(t=e^{-\beta\epsilon_{1}}\), and \(Q = e^{-i\theta}\), and where, \begin{equation} \Delta_{q,t}(\mathcal{R}) = \prod_{m=0}^{\beta-1}\prod_{i<j}\lbrack \mathcal{R}_{i}-\mathcal{R}_{j}+\beta(j-i)+m\rbrack_{q} \lbrack \mathcal{R}_{i}-\mathcal{R}_{j}+\beta(j-i)-m\rbrack_{q} \end{equation} where we have used the notation, \(\lbrack n \rbrack_{q} = q^{n/2}-q^{-n/2}\). We can now rewrite \(\Delta_{q,t}(\mathcal{R})\) in a form that clarifies the relationship to refined Chern-Simons theory, \begin{equation} \Delta_{q,t}(\mathcal{R}) =\frac{S_{0{\mathcal{R}}}S_{0{\mathcal{R}}}}{G_{\mathcal{R}}} \end{equation} where \(S_{PQ}\) is the S-matrix for refined Chern-Simons theory, but analytically continued away from integer level, and \(G_{R}\) is the finite $N$ Macdonald metric. These elements take the form, \begin{eqnarray} S_{\mathcal{R}0} = S_{00}\textrm{dim}_{q,t}(\mathcal{R}) & = & \prod_{m=0}^{\beta-1}\prod_{1\leq i<j \leq N} \lbrack \mathcal{R}_{i} - \mathcal{R}_{j} + \beta(j-i)+m \rbrack_{q} \label{eqn:qtdimbeta} \\ G_{\mathcal{R}} & = & \prod_{m=0}^{\beta-1}\prod_{1 \leq i<j \leq N}\frac{ \lbrack \mathcal{R}_{i}-\mathcal{R}_{j}+\beta(j-i)+m\rbrack_{q} }{\lbrack \mathcal{R}_{i}-\mathcal{R}_{j}+\beta(j-i)-m\rbrack_{q}}\label{eqn:metricbeta} \end{eqnarray} Putting this together, we can rewrite the entire partition function as, \begin{equation} Z_{\textrm{ref BH}}= \sum_{\mathcal{R}}\frac{(S_{\mathcal{R}0})^{2-2g}}{(G_{\mathcal{R}})^{1-g}} q^{\frac{p(\mathcal{R},\mathcal{R})}{2}}t^{p(\rho,\mathcal{R})}Q^{\sum_{i}\mathcal{R}_{i}} \end{equation} in agreement with equation \ref{eqn:qtym11}. Finally, for comparison with refined topological string theory, it is helpful to include an overall \(Q\)-independent normalization factor, \begin{equation} \alpha_{BH} = \exp\Bigg(-\frac{\epsilon_{2}^{2}}{\epsilon_{1}}\frac{\rho^{2}(p+2g-2)^{2}}{2p}+\frac{N\theta^{2}}{2p\epsilon_{1}} + \frac{\epsilon_{2}^{2}}{\epsilon_{1}}(2g-2)\rho^{2} \Bigg)\Big((t;q)_{\infty}(q;q)_{\infty}\Big)^{N(g-1)} \label{eqn:alphanorm} \end{equation} Therefore, our final result takes the form, \begin{equation} Z_{qt\textrm{YM}}(\Sigma) =\alpha_{BH} \sum_{\mathcal{R}}\frac{(S_{\mathcal{R}0})^{2-2g}}{(G_{\mathcal{R}})^{1-g}} q^{\frac{p(\mathcal{R},\mathcal{R})}{2}}t^{p(\rho,\mathcal{R})}Q^{\sum_{i}\mathcal{R}_{i}} \label{eqn:qtym} \end{equation} In conclusion, \((q,t)\)-deformed Yang-Mills gives a precise prediction for the refined black hole partition function for D4 branes wrapping an arbitrary complex line bundle over \(\Sigma\). From the discussion above, this can also be rephrased as a mathematical prediction for the \(\chi_{y}\) genus of instanton moduli space. It is important to notice that, as written, \(Z_{qt\textrm{YM}}\) is an expansion in \(q=e^{-\epsilon_{1}}\) and \(t=e^{-\epsilon_{2}}\), while the original black hole index of equation \ref{eqn:rtrace} should be expanded in \(e^{-1/\epsilon_{1}}\) and \(e^{-\gamma}\). Therefore, to extract D4/D2/D0 refined degeneracies, we must use TST duality to resum the partition function \(Z_{BH}\) so that it is written in the appropriate expansion. \subsection{Example: $\mathcal{O}(-1)\to \mathbb{P}^{1}$ \label{sec:o1}} In this section we focus on D4 branes wrapping the bundle \(\mathcal{O}(-1)\) over a genus \(g=0\) Riemann surface. $${\cal D}_0\;\; = \;\;\mathcal{O}(-1)\to\mathbb{P}^{1} $$ This geometry is simply given by blowing up \(\mathbb{C}^{2}\) at a point, which means that the corresponding instanton moduli space is especially simple. In the mathematical work of Yoshioka and Nakajima \cite{Yoshioka:1996pd, Nakajima:2003uh}, the authors proved an explicit formula for the Hodge polynomial of \(U(N)\) instantons on the blow-up geometry, \begin{eqnarray} P_{\textrm{blow-up}}(\phi_{0},\phi_{2};x,y) & = & \sum_{ch_{2},c_{1},i}e^{-\phi_{0}ch_{2}-\phi_{2} c_{1}}(-1)^{i+j}x^{i}y^{j}h_{i,j}(\mathcal{M}_{ch_{2},c_{1}}) \\ & = & \frac{1}{\eta(e^{-\phi_{0}})^{N}}\sum_{\{n_{i}\}=-\infty}^{\infty}e^{-\phi_{0}\frac{(n,n)}{2}}(xy)^{(\rho,n)}e^{-\phi_{2}\sum_{i}n_{i}} \label{eqn:yoshioka} \end{eqnarray} Note that since this geometry is toric, \(h_{i,j}\) is only nonzero if \(i=j\).\footnote{From equation \ref{eqn:hodgespin}, it follows that all D4/D2/D0 BPS states in this geometry have zero R-charge (\(R=0\)). Therefore, the ordinary and protected spin characters agree.}. Setting \(x=1\) we obtain the \(\chi_{y}\) genus of interest, \begin{equation} P_{\chi_{y}; \textrm{blow-up}}(\phi_{0},\phi_{2},y) = \frac{1}{\eta(e^{-\phi_{0}})^{N}}\sum_{\{n_{i}\}=-\infty}^{\infty}e^{-\phi_{0}\frac{(n,n)}{2}}y^{(\rho,n)}e^{-\phi_{2}\sum_{i}n_{i}} \label{eqn:yoshioka} \end{equation} We would like to compare this answer against the prediction of \((q,t)\)-deformed Yang-Mills, \begin{equation} Z_{\textrm{ref}}({\cal D}_0) = \alpha_{BH} \sum_{n_{1}\geq n_{2}\geq\cdots\geq n_{N}}q^{\frac{(n,n)}{2}}t^{(\rho,n)}Q^{\sum_{i}n_{i}}\frac{\textrm{dim}_{q,t}(\{n_{i}\})^{2}}{g(\{n_{i}\})} \end{equation} This partition function can be rewritten using the remarkable identity,\footnote{An analogue of this identity for the \(SU(N)\) case has been proven by Cherednik in \cite{Ch:MacMehta, Ch:DAHA, Ch:Diff}.} \begin{equation} \sum_{n_{1}\geq n_{2}\geq\cdots\geq n_{N}}q^{\frac{(n,n)}{2}}t^{(\rho,n)}Q^{\sum_{i}n_{i}}\frac{\textrm{dim}_{q,t}(\{n_{i}\})^{2}}{g(\{n_{i}\})} = \Bigg(\prod_{k=1}^{N-1}\prod_{j=1}^{\infty}\frac{1-q^{j}t^{k}}{1-q^{j}}\Bigg)\sum_{n_{i}=-\infty}^{\infty}q^{\frac{(n,n)}{2}}t^{(\rho,n)}Q^{\sum_{i}n_{i}} \end{equation} Since the summation on the right is over all integers \(\{n_{i}\}\), we can shift the definition of \(n\) by any integer amount. Using this freedom we can finally write the partition function as, \begin{equation} Z_{\textrm{ref}}({\cal D}_0)= \alpha_{BH}(q,t)\Bigg(\prod_{k=1}^{N-1}\prod_{j=1}^{\infty}\frac{1-q^{j}t^{k}}{1-q^{j}}\Bigg)\sum_{n_{i}=-\infty}^{\infty}q^{\frac{(n,n)}{2}}(t/q)^{(\rho,n)}Q^{\sum_{i}n_{i}} \label{eqn:blowup} \end{equation} The \(Q\)-independent prefactor in this expression is related to \(D4/D0\) degeneracies and is ambiguous because our geometry is non-compact. For this reason, in subsequent formulas we will drop it. To compare our result with the \(\chi_{y}\)-genus of instanton moduli space, we must remember that the partition function of \((q,t)\)-deformed Yang-Mills is an expansion in \(e^{-\epsilon_{k}}\), while D-brane degeneracies arise as coefficients of an expansion in \(e^{-\frac{1}{\epsilon_{1}}}\). To relate these two expansions, we can rewrite \(Z_{\textrm{ref}}\) as a product of Jacobi theta functions, \begin{equation} Z_{\textrm{ref}}({\cal D}_0)\sim \prod_{j=1}^{N}\vartheta\Big(\frac{(\epsilon_{1}-\epsilon_{2})(N+1-2j)}{4\pi i} - \frac{\theta}{2\pi},-\frac{\epsilon_{1}}{2\pi i}) \end{equation} where the Jacobi theta function is defined by \(\vartheta(z,\tau) = \sum_{n}e^{\pi i \tau n^{2} + 2\pi i n z}\). It is helpful to recall that the Jacobi theta function has the modular property, \begin{equation} \vartheta\Big(\frac{z}{\tau},-\frac{1}{\tau}\Big) = (-i\tau)^{1/2}\exp\Big(\pi i z^{2}/\tau\Big)\vartheta(z,\tau) \end{equation} Applying this transformation to the black hole partition function we obtain, \begin{equation} Z_{\textrm{ref}}({\cal D}_0)\sim \sum_{n_{i}=-\infty}^{\infty}e^{-\phi_{0} \frac{(n,n)}{2}}e^{-\phi_{2}\sum_{i}n_{i}}e^{-2\gamma(\rho,n)} \label{eqn:goodo1} \end{equation} where \(\phi_{0}\), \(\phi_{2}\), and \(\gamma\) are given by the definitions in equation \ref{eqn:cov}. As discussed above, physically this modular transformation arises from performing TST duality on the D-brane configuration. This result precisely agrees with the expected \(\chi_{y}\)-genus in equation \ref{eqn:yoshioka}. In section \ref{sec:wc}, we will give an independent physical derivation of equation \ref{eqn:yoshioka} by using the refined semi-primitive wall crossing formula. \section{Large N Factorization and the Refined OSV Conjecture \label{sec:factor}} Now that we have explained how to compute both the refined black hole partition function and the refined topological string, we would like to see how they are connected. As explained in section \ref{sec:osv}, the refined OSV conjecture predicts that the refined partition function of \(N\) D4-branes wrapping, \begin{equation} \mathcal{O}(-p) \to \Sigma_{g} \end{equation} should be equal to the square of the partition function of refined topological string on \begin{equation} \mathcal{O}(-p)\oplus\mathcal{O}(2g-2+p) \to \Sigma_{g} \end{equation} to all orders in the \(1/N\) expansion. In equations, this implies \begin{equation} Z_{qt\textrm{YM}}(\phi_{0},\phi_{2},\gamma) \sim \vert Z_{\textrm{ref top}}(\epsilon_{1},\epsilon_{2},k) \vert^{2} \end{equation} with the change of variables, \begin{eqnarray} k & = & \frac{2\pi^{2}}{\phi_{0}}\Big(\beta P + i \frac{\phi_{2}}{\pi}\Big)\nonumber \\ \epsilon_{1} & = & \frac{4\pi^{2} C}{\phi_{0}} \label{eqn:cov2} \\ \frac{\epsilon_{2}}{\epsilon_{1}} & = & \beta = 1 - \frac{\gamma}{2\pi i} \nonumber \end{eqnarray} Note that we have included the additional constant factor, \(C\). It will become clear (see equation \ref{eqn:cov}) that in our example the value for \(C\) is \(1\) so that, \begin{equation} \phi_{0} = \frac{4\pi^{2}}{\epsilon_{1}} \end{equation} This is in contrast with the more symmetric choice of \begin{equation} \phi_{0} \sim \frac{4\pi^{2}}{\epsilon_{1} + \epsilon_{2}} \end{equation} that was used in our discussion of the refined OSV conjecture in section \ref{sec:osv}. This apparent discrepancy can be resolved by postulating that \(D0\)-branes in our setup carry intrinsic charge under \((J_{3}-R)\), which will change their effective weighting in the refined partition function. As explained in section \ref{sec:n4qt}, this is expected since in the dual refined Chern-Simons construction, \(D0\)-branes carry angular momentum which causes them to be weighted by either \(q\) or \(t\), but not \(\sqrt{qt}\). We can make the change of variables in equation \ref{eqn:cov2} even more explicit for the geometries we are studying. It might seem that the D4-brane charge, \(P\), is simply the number of \(D4\) branes, \(N\), that wrap the bundle, \(C_{4} = \mathcal{O}(-p) \to \Sigma_{g}\). However, we should really measure this charge in electric \(D2\) brane units. As explained in \cite{Aganagic:2004js} these charges differ because of the nontrivial intersection number of the Riemann surface \(\Sigma\) with the four-cycle wrapped by the D4 branes, \(C_{4}\), \begin{equation} \#(\Sigma \cap C_{4}) = 2g-2+p \end{equation} leading to the identification, \begin{equation} P = N(2g-2+p) \end{equation} Now we want to test the above predictions by using the results that we have built up in sections \ref{sec:tqft} and \ref{sec:branes}, where we solved for both the refined black hole partition function and the refined topological string on these geometries. We found in section \ref{sec:branes}, that the refined brane partition function is computed by the two-dimensional \((q,t)\)-deformed Yang-Mills, whose partition function depends on two coupling constants \((\epsilon_{1},\epsilon_{2})\) and a theta-term, \(\theta\). As explained in equation \ref{eqn:cov}, these gauge theory variables are related to the refined black hole chemical potentials by, \begin{equation} \phi_{0} = \frac{4\pi^{2}}{\epsilon_{1}}, \qquad \qquad \phi_{2} = \frac{2\pi \theta}{\epsilon_{1}}, \qquad \qquad \gamma = \frac{2\pi i(\epsilon_{2}-\epsilon_{1})}{\epsilon_{1}} \end{equation} Putting this together with the predictions of the refined OSV conjecture, we find that the large \(N\) limit of \((q,t)\)-deformed Yang-Mills should be equal to the square of the refined topological string with the K\"ahler modulus equal to, \begin{equation} k = 2\pi i \frac{X_{1}}{X_{0}} = \frac{1}{2}(2g-2+p)N\epsilon_{2}+i\theta \label{eqn:kahler} \end{equation} and the topological string couplings \((\epsilon_{1},\epsilon_{2})\) identified with the same variables in the Yang-Mills theory. \begin{figure}[htp] \centering \includegraphics[scale=0.80]{CompositeTableau.pdf} \caption{The composite Young Tableau, \(R_{+}\overline{R}_{-}\), is shown for two \(SU(N)\) representations $R_{+}$ and $R_{-}$. \label{fig:composite}} \end{figure} To test this precise prediction, we must carefully study the large \(N\) limit of \((q,t)\)-deformed Yang-Mills. The key observation, first made for ordinary two-dimensional Yang-Mills in \cite{Gross:1993hu}, is that representations of \(SU(N)\) for large \(N\) can be viewed as composites of two Young Tableaux as shown in Figure \ref{fig:composite}. This splitting can be thought of as a splitting of the Hilbert space as \(N\to \infty\) into two chiral pieces, \begin{equation} \mathcal{H} \sim \mathcal{H}_{+} \otimes \overline{\mathcal{H}}_{-} \end{equation} If the dynamics of the theory respect this splitting, then the partition function will also factorize into two pieces as predicted by the refined OSV conjecture. In this section we will show that the partition function does indeed satisfy this factorization, after properly accounting for the sum over \(RR\) fluxes and asymptotic boundary conditions. It is important to note that this splitting is valid to all orders in \(1/N\), but it is not valid nonperturbatively since the two Young Tableaux will interfere with each other at finite \(N\). It would be interesting to compute the nonperturbative refined corrections to our result \cite{nonpertme}. As in the unrefined case, it is helpful to split our discussion into the genus \(g\geq 1\) and the genus \(g=0\) cases. We begin with the higher genus geometries. \subsection{Genus $g\geq 1$ Case} As explained in equation \ref{eqn:qtym}, the refined black hole partition function for \(N\) \(D4\) branes wrapping \(\mathcal{O}(-p) \to \Sigma_{g}\) is given by, \begin{equation} Z_{qt\textrm{YM}}(\Sigma_{g},p) =\alpha_{BH} \sum_{\mathcal{R}}\Bigg(\frac{(S_{\mathcal{R}0})^2}{G_{\mathcal{R}}}\Bigg)^{1-g} q^{\frac{p(\mathcal{R},\mathcal{R})}{2}}t^{p(\rho,\mathcal{R})}Q^{\sum_{i}\mathcal{R}_{i}} \end{equation} The sum is over all \(U(N)\) representations which we denote by \(\mathcal{R}\). The normalization constant \(\alpha_{\textrm{BH}}\) defined in equation \ref{eqn:alphanorm} is included for convenience when making the connection with refined topological strings. When we take the large \(N\) limit, it will be helpful to decompose each \(U(N)\) representation into an \(SU(N)\) and \(U(1)\) representation. Recall that representations of \(U(N)\) are labeled by integers, \(\mathcal{R}_{i}\), such that \(\mathcal{R}_{1}\geq \mathcal{R}_{2} \geq \cdots \geq \mathcal{R}_{N}\) where the \(\mathcal{R}_{i}\) can take positive or negative values. We can decompose this into a representation of \(SU(N)\) and a representation of \(U(1)\) by rewriting it as, \begin{eqnarray} \mathcal{R}_{i} & = & R_{i} + r \;\;\qquad i=1,\ldots,N-1 \\ \mathcal{R}_{N} & = & r \nonumber \end{eqnarray} where the \(R_{i}\) label an \(SU(N)\) representation. Then the \(U(1)\) charge is given by \begin{equation} m= \vert R \vert + Nr \end{equation} where \(r \in \mathbb{Z}\). We can rewrite the terms appearing in the partition function, as, \begin{eqnarray} \vert \mathcal{R} \vert & = & \sum_{k}\mathcal{R}_{k} = m \nonumber \\ (\mathcal{R}, \mathcal{R}) & = & \sum_{k}\mathcal{R}_{k}^{2} = \sum_{k}R_{k}^{2} + \frac{m^{2}}{N} - \frac{\vert R \vert^{2}}{N} \label{eqn:undecomp} \\ 2(\rho, \mathcal{R}) & = & \sum_{k}\mathcal{R}_{k}(N+1-2k) = \sum_{k}R_{k}(1-2k)+N\vert R \vert \nonumber \end{eqnarray} For genus \(g\neq 1\), the partition function also involves the metric and \((q,t)\)-dimension. These quantities are the same for the \(U(N)\) representation \(\mathcal{R}\) and its \(SU(N)\) part \(R\), \begin{eqnarray} \textrm{dim}_{q,t}(\mathcal{R}) & = & \textrm{dim}_{q,t}(R) \\ g_{\mathcal{R}} & = & g_{R} \nonumber \end{eqnarray} From equations \ref{eqn:metricbeta} and \ref{eqn:qtdimbeta}, this is true because both quantities can be written as functions only of the differences \(\mathcal{R}_{i}-\mathcal{R}_{j}\) which are independent of the \(U(1)\) charge. Now we would like to study the large \(N\) limit of this theory. As mentioned above, in this limit each \(SU(N)\) representation can be decomposed into a composite of two representations as depicted in Figure \ref{fig:composite}. If we form the composite representation of \(R_{+}\) and \(R_{-}\), which we denote by \(R_{+}\overline{R}_{-}\), then its row lengths are given by, \begin{equation} (R_{+}\overline{R}_{-})_{i} = (R_{-})_{1}+(R_{+})_{i}-(R_{-})_{N+1-i} \end{equation} Note that for this to be a good composite representation the two representations should not interact with each other, so that either \((R_{+})_{i}\) or \((R_{-})_{N+1-i}\) is equal to zero for each \(i\). As explained above, we can neglect these interactions provided that we only study the perturbative expansion in \(1/N\). Now we want to study how the quantities in equation \ref{eqn:undecomp} decompose for a \(U(N)\) representation \(\mathcal{R}\) whose \(SU(N)\) part consists of the composite representation \(R_{+}\overline{R}_{-}\). In this case the \(U(1)\) charge is given by, \(m = Nl + \vert R_{+}\vert - \vert R_{-} \vert\), where we have defined \(l=r+(R_{-})_{1}\). Then we find the decomposition, \begin{eqnarray} \vert \mathcal{R} \vert & = & \sum_{i}\mathcal{R}_{i} = \vert R_{+}\vert - \vert R_{-} \vert +Nl \nonumber \\ (\mathcal{R},\mathcal{R}) & = & \sum_{i}\mathcal{R}_{i}^{2} = \vert \vert R_{+} \vert \vert^{2} + \vert \vert R_{-} \vert \vert^{2} + Nl^{2} + 2l(\vert R_{+}\vert - \vert R_{-} \vert) \\ 2(\rho, \mathcal{R}) & = & \sum_{k}\mathcal{R}_{k}(N+1-2k) = -\vert \vert R_{+}^{T} \vert \vert^{2} - \vert \vert R_{-}^{T} \vert \vert^{2} + N\vert R_{+} \vert + N \vert R_{-} \vert \nonumber \end{eqnarray} where \(\vert \vert R \vert\vert^{2} = \sum_{k}R_{k}^{2}\) and \(\vert \vert R^{T} \vert \vert^{2} = \sum_{k}(2k-1)R_{k}\). Putting together these results, we can rewrite the refined black hole partition function as, \begin{eqnarray} Z_{\textrm{ref BH}} & = & \alpha_{\textrm{BH}} \sum_{l\in \mathbb{Z}}\sum_{R_{+},R_{-}}\Bigg(\frac{G_{R_{+}\overline{R}_{-}}}{\big(S_{0R_{+}\overline{R}_{-}}\big)^{2}}\Bigg)^{g-1}q^{\frac{p}{2}\big(\vert\vert R_{+}\vert\vert^{2} + \vert\vert R_{-} \vert \vert^{2}\big)} \\ & & t^{-\frac{p}{2}\big(\vert\vert R^{T}_{+}\vert\vert^{2} + \vert\vert R^{T}_{-} \vert \vert^{2}\big)} e^{-i\theta(\vert R_{+} \vert - \vert R_{-} \vert)}q^{lp(\vert R_{+} \vert - \vert R_{-} \vert)}t^{\frac{Np}{2}(\vert R_{-} \vert + \vert R_{+} \vert)}q^{\frac{N p l^{2}}{2}} e^{-i\theta N l} \nonumber \end{eqnarray} With the refined OSV relation in mind, we can use the formula for the K\"ahler modulus in equation \ref{eqn:kahler} to rewrite \(\alpha_{\textrm{BH}}\) as, \begin{eqnarray} \alpha_{\textrm{BH}} & = & \Big((t;q)_{\infty}(q;q)_{\infty}\Big)^{N(g-1)} \\ & & \cdot \exp\Bigg(-\frac{1}{\epsilon_{1}\epsilon_{2}}\frac{k^{3}+\overline{k}^{3}}{6p(p+2g-2) }+ \beta\frac{(k+\overline{k})(p+2g-2)}{24p} + \epsilon_{1}\beta^{2}\rho^{2}(2g-2) \Bigg) \nonumber \end{eqnarray} So far we have explained how the framing factor and \(\theta\)-dependent terms factorize, but we still need to understand the metric and \((q,t)\)-dimension. Their factorization properties are derived in Appendix \ref{sec:qtfac}, with the result that, \begin{eqnarray} \frac{G_{R_{+}\overline{R}_{-}}}{q^{2\beta^{2}\rho^{2}} \big(S_{0R_{+}\overline{R}_{-}}\big)^{2}} & = &\Big(M(q,t)M(t,q)\Big)^{-1}\Big((t;q)_{\infty}(q;q)_{\infty}\Big)^{-N}T_{R_{+}}^{2}T_{R_{-}}^{2} Q^{\vert R_{+}\vert + \vert R_{-} \vert} \nonumber \\ & & \cdot \frac{g_{R_{+}}g_{R_{-}}K_{R_{+}R_{-}}(Q\frac{q}{t})K_{R_{+}R_{-}}(Q)}{W_{R_{+}}^{4}W_{R_{-}}^{4}} \end{eqnarray} where we have defined \(Q = t^{N}\) and used the framing factor, \(T_{R} = q^{\vert \vert R \vert \vert^{2}/2}t^{-\vert\vert R^{T} \vert \vert^{2}/2}\), and where \begin{equation} K_{R_{+}R_{-}}(Q;q,t) = \sum_{P} \frac{1}{g_{P}}Q^{\vert P \vert}(t/q)^{\vert P \vert}W_{PR_{+}}(q,t)W_{PR_{-}}(q,t) \end{equation} From these factorization formulas, we can write the entire refined black hole partition function as a sum over chiral blocks, \begin{equation} Z_{\textrm{ref BH}} = \sum_{l \in \mathbb{Z}}\sum_{R_{1},\cdots R_{2g-2}}Z^{+}_{R_{1},\cdots, R_{2g-2}}(k+pl\epsilon_{1}) Z^{+}_{R_{1},\cdots, R_{2g-2}}(\overline{k}-pl\epsilon_{1}) \label{eqn:largenosv} \end{equation} where the chiral block is defined by, \begin{eqnarray} Z^{+}_{R_{1},\cdots, R_{2g-2}}(k) & = & Z_{0}(q,t) \cdot \big(t^{\frac{N}{2}}\big)^{\vert R_{1}\vert + \cdots \vert R_{g-1}\vert}\big(t^{\frac{N+1}{2}}q^{-\frac{1}{2}}\big)^{\vert R_{g}\vert + \cdots \vert R_{2g-2}\vert} \\ & & \sum_{R}(T_{R})^{p+2g-2}e^{-k\vert R\vert}\frac{W_{R_{1}R}(q,t) \cdots W_{R_{2g-2}R}(q,t)}{W_{0R}^{4g-4}}\frac{(g_{R})^{g-1}}{g_{R_{1}}\cdots g_{R_{2g-2}}} \nonumber \end{eqnarray} where \(k=\frac{1}{2}(p+2g-2)N\epsilon_{2}+i\theta\) and where we used the definition of the degree zero piece, \(Z_{0}\), defined in equation \ref{eqn:degzero}. Now we come to the physical interpretation of these chiral blocks. First notice that the chiral block is precisely the refined topological string amplitude for the geometry \begin{equation} \mathcal{O}(2g-2+p)\oplus \mathcal{O}(-p) \to \Sigma_{g} \end{equation} with branes in the fibers over \(2g-2\) points, as explained in section \ref{sec:fiberbranes}, \begin{equation} Z^{+}_{R_{1},\cdots, R_{2g-2}}(k) = Z_{\textrm{ref top, } R_{1},\cdots, R_{2g-2}}(k) \end{equation} The factors, \(t^{\frac{1}{2}N\vert R_{i} \vert}\) appear because these branes have been moved in the fiber away from the origin. These ``ghost branes'' were explained in \cite{Aganagic:2005dh} as parametrizing noncompact K\"ahler moduli in the fiber directions. Naively, one might expect there to be noncompact moduli over every point on the Riemann surface. However, these moduli can be localized by using the symmetries corresponding to meromorphic vector fields on \(\Sigma\). Since these vector fields have \(2g-2\) poles generically, we find ghost branes over precisely \(2g-2\) points. This picture must be modified slightly in the refined case, since not all of these branes are the same. They split into two groups of \(g-1\) branes, which differ only by their fiber K\"ahler parameters which are either \begin{equation} k_{f} = \frac{1}{2}N\epsilon_{2}, \qquad \textrm{or} \qquad k_{f} = \frac{1}{2}N\epsilon_{2} + \frac{1}{2}(\epsilon_{2}-\epsilon_{1}) \end{equation} It would be interesting to derive this splitting from first principles. Another important aspect of our formula for large N factorization is the sum over the \(U(1)\) charge, \(l\). As in the unrefined case \cite{Vafa:2004qa}, we interpret this as arising from the sum over \(RR\)-flux through the Riemann surface. Alternatively, this sum arises because the black hole partition function is trivially invariant under shifts \(\phi_{2} \to \phi_{2} + 2\pi i p n\).\footnote{The extra factor of \(p\) arises because the black holes have charges, \(Q_{2} \in \frac{1}{p}\mathbb{Z}\), as explained in \cite{Aganagic:2004js}.} The sum over \(U(1)\) charge enforces this same periodicity on the topological string side of the correspondence. \subsection{Genus $g=0$ Case \label{sec:genuszero}} The genus zero case works much the same way as the higher genus case. By decomposing the \(U(N)\) representations into \(U(1)\) and composite \(SU(N)\) representations, we can write the corresponding brane partition function as, \begin{eqnarray} Z_{\textrm{ref BH}} & = & \alpha \sum_{l\in \mathbb{Z}}\sum_{R_{+},R_{-}} \frac{(S_{0R_{+}\overline{R}_{-}})^2}{G_{R_{+}\overline{R}_{-}}} q^{\frac{p}{2}\big(\vert\vert R_{+}\vert\vert^{2} + \vert\vert R_{-} \vert \vert^{2}\big)}t^{-\frac{p}{2}\big(\vert\vert R^{T}_{+}\vert\vert^{2} + \vert\vert R^{T}_{-} \vert \vert^{2}\big)} \nonumber \\ & & e^{-i\theta(\vert R_{+} \vert - \vert R_{-} \vert)}q^{lp(\vert R_{+} \vert - \vert R_{-} \vert)}t^{\frac{Np}{2}(\vert R_{-} \vert + \vert R_{+} \vert)}q^{\frac{N p l^{2}}{2}} e^{-i\theta N l} \end{eqnarray} However, since \(\textrm{dim}_{q,t}(R_{+}\overline{R}_{-})\) appears in the numerator, we will use the other set of identities from appendix \ref{sec:qtfac} to give, \begin{eqnarray} \frac{q^{2\beta^{2}\rho^{2}} (S_{0R_{+}\overline{R}_{-}})^2}{G_{R_{+}\overline{R}_{-}}} & = & M(q,t)M(t,q) \Big((t;q)_{\infty}(q;q)_{\infty}\Big)^{N}T_{R_{+}}^{-2}T_{R_{-}}^{-2} Q^{-\vert R_{+}\vert - \vert R_{-} \vert} \nonumber \\ & & \cdot \frac{N_{R_{+}R_{-}}(Q\frac{q}{t})N_{R_{+}R_{-}}(Q)}{g_{R_{+}}g_{R_{-}}} \end{eqnarray} where, \begin{equation} N_{R_{+}R_{-}}(Q,q,t) := \sum_{P}\frac{1}{g_{P}}Q^{\vert P\vert}(t/q)^{\vert P\vert}\widetilde{W}_{RP}(q,t)W_{PS}(q,t) \end{equation} and where \(\widetilde{W}_{RP}\) is the cap amplitude for placing a brane in the base and an anti-brane in the fiber as explained in section \ref{sec:fiberbranes}. Therefore, we can rewrite the black hole partition function as a sum over chiral blocks, \begin{equation} Z_{\textrm{ref BH}} = \sum_{l \in \mathbb{Z}}\sum_{R_{1}, R_{2}}Z^{+}_{R_{1},R_{2}}(k+pl\epsilon_{1}) Z^{-}_{R_{1}, R_{2}}(\overline{k}-pl\epsilon_{1}) \label{eqn:largenosvg0} \end{equation} where the K\"ahler parameter is given by, \begin{equation} k = \frac{1}{2}(p-2)N\epsilon_{2}+i\theta \end{equation} It is important to notice that in the genus zero case, the chiral and anti-chiral blocks are not precisely the same. The chiral block is equal to, \begin{equation} Z^{+}_{R_{1},R_{2}}(k) = Z_{0}(q,t) t^{\frac{1}{2}N\vert R_{1}\vert}\big(t^{\frac{N+1}{2}}q^{-\frac{1}{2}}\big)^{\vert R_{2} \vert} \sum_{R}(T_{R})^{p-2}e^{-k\vert R \vert}\frac{W_{R_{+}R_{1}}(q,t) W_{R_{+}R_{2}}(q,t)}{g_{R}\; g_{R_{1}}\, g_{R_{2}}} \end{equation} while the anti-chiral block takes the slightly different form, \begin{equation} Z^{-}_{R_{1},R_{2}}(k) = Z_{0}(q,t) t^{\frac{1}{2}N\vert R_{1}\vert}\big(t^{\frac{N+1}{2}}q^{-\frac{1}{2}}\big)^{\vert R_{2} \vert} \sum_{R}(T_{R})^{p-2}e^{-k\vert R \vert}\frac{\widetilde{W}_{R_{+}R_{1}}(q,t) \widetilde{W}_{R_{+}R_{2}}(q,t)}{g_{R}\; g_{R_{1}}\, g_{R_{2}}} \end{equation} As in the higher genus case, the refined ghost branes split into two types depending on their fiber K\"ahler moduli and we obtain a sum over RR-flux. The chiral block computes precisely the refined topological string amplitude on \(\mathcal{O}(p-2)\oplus \mathcal{O}(-p) \to \mathbb{P}^{1}\) with two ``ghost'' branes in the fiber. The main difference between the higher genus and genus zero cases is that here the anti-chiral amplitude is the refined topological string on the same geometry but with two anti-branes rather than branes in the fiber. We can use the same argument as before for why there are precisely two ghost branes. However, now localization by using a generic vector field on the \(\mathbb{P}^{1}\) will have two zeros rather than poles. \section{Black Hole Entropy and Refined Wall Crossing \label{sec:wc}} In section \ref{sec:o1}, we used \((q,t)\)-deformed Yang-Mills to compute the entropy of D-branes wrapping the blow-up geometry \(\mathcal{O}(-1) \to \mathbb{P}^{1}\) and found agreement with the mathematical result of \cite{Yoshioka:1996pd}. In this section, we explain an alternative way to compute this black hole partition function by using refined wall-crossing. We follow the approach of \cite{Nishinaka:2010qk, Nishinaka:2010fh}, giving a refined generalization of their unrefined computations. \begin{figure}[htp] \centering \includegraphics[scale=1.15]{ConifoldFlop.pdf} \caption{The flop transition for the conifold. $N$ D4-branes wrap the shaded four-cycle, which changes topology under the flop. \label{fig:conifoldflop}} \end{figure} We start by studying the resolved conifold, \(\mathcal{O}(-1)\oplus \mathcal{O}(-1) \to \mathbb{P}^{1}\), with \(N\) D4-branes wrapping the four cycle \(C_{4} = \mathcal{O}(-1)\to \mathbb{P}^{1}\). We would like to compute the refined degeneracies of these \(D4\) branes bound to lower dimensional branes. Note that since \(C_{4}\) contains the compact two-cycle, \(\mathbb{P}^{1}\), D2 branes can form bound states with the stack of \(D4\)s. The key insight of \cite{Nishinaka:2010qk} is that by varying the K\"ahler modulus of the \(\mathbb{P}^{1}\), which we denote by \(z\), the refined partition function will simplify in certain chambers. Specifically, by sending \(z \to -\infty\), the conifold will undergo a flop transition, and the four-cycle, \(C_{4}\) will become a cycle wrapping only the fiber directions, as in shown in Figure \ref{fig:conifoldflop}. Most importantly, this new four-cycle is topologically \(\mathbb{C}^{2}\) and does not have any compact two-cycles. Thus, in this chamber, only D4 and D0 branes can bind, and the refined partition function simplifies dramatically. Now by starting in this simple chamber, we can vary the K\"ahler parameter \(z\) and keep track of how the refined partition function jumps. The partition function is locally constant, but along real codimension-one walls of marginal stability it will jump. Since the conifold geometry is fairly simple, we can identify all walls of marginal stability and jumps as we take \(z\) from \(-\infty\) to \(\infty\). This will allow us to explicitly compute the refined partition function at \(z=\infty\) and find agreement with our \((q,t)\)-deformed Yang-Mills computation in section \ref{sec:o1}. To find these walls, we must first compute the central charge of BPS bound states. Note that since our geometry is noncompact, the central charge of the \(D4\) branes is infinite and its phase is not well-defined. As explained originally in \cite{Jafferis:2008uf}, this can be remedied by starting with a compact geometry and including a component of the complexified K\"ahler form along the direction that is becoming noncompact. The result is that the central charge of the \(D4\) brane is given by, \begin{equation} Z(D4) = - \frac{1}{2}\Lambda^{2}e^{2i\phi} \end{equation} where \(\Lambda \gg 1\), and ultimately we want to take \(\Lambda \to \infty\) to obtain the resolved conifold. Note that \(\phi\) is still a free real parameter, but for our purposes we will keep it fixed and only vary the complexified K\"ahler parameter, \(z\). Now we can consider a more general D4/D2/D0 bound state with charges, \begin{equation} \Gamma = (P_{6},P_{4},Q_{2},Q_{0}) = (0,N,m,n) \end{equation} Its central charge is given by, \begin{equation} Z(\Gamma) = - \frac{1}{2}N\Lambda^{2}e^{2i\phi}+mz+n \end{equation} This state can decay as \(\Gamma \to \Gamma_{1} + \Gamma_{2}\) precisely when \(Z(\Gamma_{1})\) and \(Z(\Gamma_{2})\) are aligned. Depending on the charges of \(\Gamma_{i}\), there are two types of walls that we must consider. First, \(\Gamma\) could decay into two states that each have nonzero \(D4\)-brane charge. Such fragmentation of the stack of D4 branes was discussed in this context in \cite{Nishinaka:2010fh}. Although the alignment of central charges seems to suggest that these fragments can form, we argue that fragmentation walls are not physical in our setup. The crucial point is that the \(D4\) branes remain noncompact throughout moduli space. From a field theory perspective fragmentation corresponds to changing a scalar field's vev. But for a field theory on noncompact spacetime, this vev is a background parameter of the theory and changing it would cost infinite energy. Therefore, we conclude that because of the noncompactness of the \(D4\)-branes, there can be no binding or decay across these walls, and they can be consistently ignored. The second type of wall is much more interesting for us and involves the decay \(\Gamma \to \Gamma_{1} + m\Gamma_{2}\) where \(\Gamma_{2}\) only has D2/D0 charge. From the Gopakumar-Vafa invariants of the conifold, it follows that the only BPS D2/D0 bound states are, \begin{eqnarray} & & \Omega_{0}(0,0,0,n;y) = -2 \nonumber \\ & & \Omega_{0}(0,0,\pm 1,n;y) = 1 \end{eqnarray} where \(\Omega_{j}(\Gamma)\) is the refined degeneracy for states with spin \(j\). Now we would like to find the walls of marginal stability where \(\Gamma = (0,N,m,n)\) can decay into these bound states. Since \(\Lambda \gg 1\), the phase of the central charge is equal to \(\arg(Z(\Gamma)) = 2\phi+ \pi\), which means that the walls of marginal stability for \(\Gamma\) will be independent of \(N\), \(m\), and \(n\), provided \(N>0\). First we can consider the pure \(D0\)-brane decay channel where \(\Gamma_{2}=(0,0,0,n)\). However, the central charge of \(\Gamma_{2}\) is always real, which means that for a generic fixed choice of \(\phi\), the central charges of \(\Gamma\) and \(\Gamma_{2}\) will never align. This means that there are no walls of marginal stability associated with \(D0\) decay. Next, we consider the second possibility of a decay involving \(\Gamma_{2}=(0,0,\pm 1,n)\). In this case, the phase of the central charge is given by, \(\arg(Z(\Gamma_{2})) = \arg(\pm z + n)\), which implies that a wall of marginal stability can occur as we vary \(z\). We will denote these walls of marginal stability by \(W^{\pm 1}_{n}\), and they are given explicitly by, \begin{eqnarray} W^{1}_{n}:\qquad \phi & = & \frac{1}{2}\arg(-z-n) \\ W^{-1}_{n}:\qquad \phi & = & \frac{1}{2}\arg(z-n) \end{eqnarray} These walls of marginal stability are shown in Figure \ref{fig:conifoldwalls}. \begin{figure}[htp] \centering \includegraphics[scale=1.25]{ConifoldWalls2.pdf} \caption{Walls of marginal stability for the conifold moduli space. The path we follow is shown by the arrow. \label{fig:conifoldwalls}} \end{figure} Now that we have identified all walls of marginal stability, we want to study how the partition function jumps across these walls. For general decays, this jump can be quite complicated and is determined by the formula of Kontsevich and Soibelman \cite{KSoriginal}. However, for our purposes we only must deal with semi-primitive decays of the form, \(\Gamma \to \Gamma_{1} + m\Gamma_{2}\). The refined semi-primitive wall-crossing formula was computed in \cite{Dimofte:2009bv}, and is given by, \begin{equation} \sum_{n=0}^{\infty}\Omega(\Gamma_{1}+n\Gamma_{2};y)x^{n} = \Omega(\Gamma_{1};y)\prod_{k=1}^{\infty}\prod_{j=1}^{k\vert \langle \Gamma_{1},\Gamma_{2} \rangle \vert}\prod_{n}\Big(1+(-1)^{n}x^{k}y^{\frac{k\vert \langle \Gamma_{1},\Gamma_{2} \rangle \vert +1 - 2j}{2}}\Big)^{(-1)^{n}\Omega_{n}(k\Gamma_{2})} \end{equation} where we have used the intersection form \(\langle \Gamma_{1},\Gamma_{2}\rangle\). In the mirror IIB geometry, this intersection form is simply the geometric intersection number of the corresponding lagrangian three-cycles. In IIA, it is equal to, \begin{equation} \langle \Gamma, \Gamma' \rangle = P_{6}\cdot Q'_{0}-P'_{6}\cdot Q_{0}+P_{4}\cdot Q'_{2}-P'_{4}\cdot Q_{2} \end{equation} Finally, in the exponent of the semi-primitive wall-crossing formula we have used the refined degeneracies, \(\Omega_{n}(\Gamma)\), that compute the index of states with spin \(n\). Now we that we have explained the refined semi-primitive wall-crossing formula, we want to apply this to our setup. We start with some BPS state \(\Gamma = (0,N,m,n)\). At the \(W^{1}_{n}\) wall, our state will decay into \(\Gamma_{1}+\Gamma_{2}\) where \(\Gamma_{2}=(0,0,1,n)\). This means that the intersection is given by \(\vert \langle \Gamma_{1},\Gamma_{2} \rangle \vert = N\). From our discussion above, the semi-primitive wall-crossing formula simplifies since \(\Omega(k\Gamma_{2})=0\) for \(k>1\), and \(\Omega_{n}(\Gamma_{2}) = \delta_{0,n}\). Putting this together, we can write the refined black hole partition function using chemical potentials \(\tilde{q}\) and \(\widetilde{Q}\) for the D0 and D2-brane charge respectively. Then we find that the partition function jumps across the wall \(W^{1}_{n}\) as, \begin{equation} Z_{\textrm{ref BH}}(\widetilde{Q},\tilde{q},y;z) \to Z_{\textrm{ref BH}}(\widetilde{Q},\tilde{q},y;z)\prod_{j=1}^{N}\Big(1+y^{\frac{N+1}{2}-j}\widetilde{Q} \tilde{q}^{n}\Big) \end{equation} Note that the wall-crossing factor takes the form of the Fock space character for a spin \( \frac{N-1}{2} \) multiplet. Similarly, crossing the wall \(W^{-1}_{n}\) results in the jump, \begin{equation} Z_{\textrm{ref BH}}(\widetilde{Q},\tilde{q},y;z) \to Z_{\textrm{ref BH}}(\widetilde{Q},\tilde{q},y;z)\prod_{j=1}^{N}\Big(1 + y^{\frac{N+1}{2}-j}\widetilde{Q}^{-1}\tilde{q}^{n}\Big) \end{equation} Now having understood how to cross individual walls, we want to follow a path in moduli space that connects the flopped geometry to the the large volume limit of interest. As shown in Figure \ref{fig:conifoldwalls}, we can take \(z=\frac{1}{2} + ir\) and follow the path from \(r=-\infty\) to \(r=+\infty\). This path crosses all \(W^{1}_{n}\) and \(W^{-1}_{n}\) walls for \(n>0\), along with the wall \(W^{1}_{0}\). This implies the relationship, \begin{equation} Z_{+\infty}(\tilde{q},\widetilde{Q},y)= Z_{-\infty}(\tilde{q},y) \prod_{j=1}^{N}\Bigg\{ \big(1 + y^{\frac{N+1}{2}-j}\widetilde{Q}\big)\prod_{n=1}^{\infty} \big(1 + y^{\frac{N+1}{2}-j}\widetilde{Q}\tilde{q}^{n}\big)\big(1 + y^{j-\frac{N+1}{2}}\widetilde{Q}^{-1}\tilde{q}^{n}\big) \Bigg\} \end{equation} Using the Jacobi Triple product, this can be rewritten as, \begin{eqnarray} Z_{+\infty}(\tilde{q},\widetilde{Q},y) & = & Z_{-\infty}(\tilde{q},y) \Bigg(\prod_{n=1}^{\infty}(1-\tilde{q}^{n})^{-N}\Bigg)\sum_{\{n_{i}\}}\tilde{q}^{\frac{1}{2}(n,n)}(\widetilde{Q}\tilde{q}^{-1/2})^{\sum_{i}n_{i}} y^{\sum_{i}(\frac{N+1}{2}-i)n_{i}} \\ & \sim & \sum_{\{n_{i}\}}e^{-\phi_{0}\frac{1}{2}(n,n)}e^{-\phi_{2}\sum_{i}n_{i}} y^{\sum_{i}(\frac{N+1}{2}-i)n_{i}} \end{eqnarray} where we have identified \(e^{-\phi_{0}} = \tilde{q}\) and \(e^{-\phi_{2}} = \widetilde{Q}\tilde{q}^{-1/2}\).\footnote{This shift in charges arises because of the Freed-Witten anomaly \cite{Freed:1999vc}, which implies that the spacetime D-brane charge, \(\Gamma\), is related to the chern character of a vector bundle, \(E\), over a four-cycle \(S\), by \(\Gamma = \textrm{ch}(E) e^{\frac{1}{2}c_{1}(S)}\). This leads to the above nontrivial relationship between the D-brane charges seen by wall-crossing, and the instanton charges seen by \((q,t)\)-deformed Yang-Mills.} Up to the D0/D4-brane bound states which are determined by the \(z=-\infty\) chamber, this formula agrees precisely with the partition function computed in equation \ref{eqn:goodo1}. Thus, we have given two independent derivations of this mathematical formula, from gauge theory and from wall-crossing. \acknowledgments{We thank Chris Beem, Emanuel Diaconescu, Tudor Dimofte, Ori Ganor, Lotte Hollands, Andrei Okounkov, Vasily Pestun, Ashoke Sen, Andy Strominger, Cumrun Vafa, and Kevin Wray for discussions. We also thank the 2012 Simons Workshop in Mathematics and Physics for hospitality while this work was being completed. The research of K.S. is supported by the Berkeley Center for Theoretical Physics and by the National Science Foundation (award number 0855653). The research of M.A. is supported in part by the Berkeley Center for Theoretical Physics, by the National Science Foundation (award number 0855653), by the Institute for the Physics and Mathematics of the Universe, and by the US Department of Energy under Contract DE-AC02-05CH11231. }
\section{Introduction} Natural images often display various forms of anisotropy. For a wide range of applications, anisotropy has been quantified through regularity characteristics and features that strongly differ when measured in different directions. This is, for instance, the case in medical imaging (osteoporosis, muscular tissues, mammographies,...), cf. e.g.~\cite{bonami:estrade:2003,bierme:meerschaert:scheffler:2009}, hydrology~\cite{ponson:2006}, fracture surfaces analysis~\cite{davies:hall:1999},\ldots. For such images, a key issue consists first in describing, within a suitable framework, the anisotropy of the texture, and then in defining regularity anisotropy parameters that can actually and efficiently be measured via numerical procedures and further involved into e.g., classification schemes. This requires the design of a mathematical framework that allows to define and estimate these parameters. Such a program can be split into several questions, some of them having already been either solved or, at least, patially addressed. A first issue lies in introducing global and local notions of anisotropic regularity, which emcompass and extend (isotropic) regularity spaces, such as Sobolev or Besov spaces, and the classical notion of pointwise H\"older regularity. To model anisotropy, the particular setting of an anisotropic self--similar field driven by two parameters (an {\it anisotropy matrix} and a {\it self--similarity index}) has been introduced and studied in~\cite{bierme:meerschaert:scheffler:2009}, where it is used as a relevant model to describe osteoporosis. In~\cite{clausel:vedel:2010}, the question of defining in a proper way the concept of anisotropy of an image in relation to its global regularity has been addressed. It has notably been shown that these two parameters can be recovered without {\it a--priori knowledge} of the characteristics of the model, by studying the global smoothness properties of the process. Furthermore, some of the properties characterizing anisotropy are revealed by the regularity of the sample paths when analyzed with functional spaces well-adapted to anisotropy: Anisotropic Besov spaces. This preliminary study thus showed the central role that such spaces should play in the mathematical modeling of random anisotropic textures. The introduction of these spaces traces back to the study of some PDEs, cf. e.g.~\cite{triebel:1978}, for the study of semi-elliptic pseudo-differential operators whose symbols have different degrees of smoothness along different directions, the reader is also referred to~\cite{aimar:gomez:2012}, and references therein, for a recent use of such spaces for optimal regularity results for the heat equation. Other types of directional function spaces have also been considered, cf. e.g.~\cite{bownik:2003} for the variant supplied by Hardy directional spaces. A second crucial issue consists in obtaining a simple characterization of these spaces on a ``dictionary''. Indeed, the challenge is to measure the critical exponent of any image in anisotropic Besov spaces for different anisotropies using the same analysis tool. Wavelet analysis is well--known to be an efficient tool for measuring smoothness in a large range of functional spaces (cf.~\cite{meyer:1990} for details). Here, however, the main point is that the anisotropy of the analyzing spaces must not be set a priori to a fixed value but instead be allowed to vary. Specific bases are thus looked for, which would serve as a common dictionary for anisotropic Besov spaces with different anisotropies. It is natural that one should use some form of anisotropic wavelets such as curvelets, bandelets, contourlets, shearlets, ridgelets, or wedgelets (see e.g.~\cite{jacques:duval:chaux:peyre:2012} for an thorough review of these representation systems and a comparison of their properties for image processing); natural criteria of choice being, on the mathematical side, that these variations on isotropic wavelets supply bases for the corresponding anisotropic spaces, and, on the applied side, that practically tractable procedure can be devised and implemented to permit the characterization of real-world data according to these function spaces. Many authors addressed this problem, and proposed different solutions, depending on the precise definition of anisotropic space they started with, as well as on the anisotropic basis they used, cf. e.g.~\cite{devore:konyagin:temlyakov:1998,hochmuth:2002a,hochmuth:2002b,long:triebel:1979,triebel:2004} and also the recent papers~\cite{garrigos:tabacco:2002} by G. Garrig\'os and A. Tabacco, and~\cite{haroske:tamasi:2005} by D. Haroske and E. Tam\'asi, which contain numerous references on the subject. Note also that, in several cases, a particular type of anisotropy was considered: Parabolic anisotropy (where a contraction by $\lambda$ in one direction is associated with a contraction by $\lambda^2$ along the orthogonal direction, \cite{lakhonchai:sampo:sumetkijakan:2012,sampo:sumetkijakan:2009,nualtong:sumetkijakan:2005}, and references therein, in particular for applications to directional regularity), the corresponding dictionaries being in that case curvelets or contourlets (corresponding to the Hart--Smith decomposition in the continuous setting, see~\cite{smith:1998}). This particular choice of anisotropy was motivated from application to PDEs (see e.g., \cite{guo:labate:2008,candes:demanet:2005} where curvelets and ridgelets are used for the study of Fourier integral operators, with applications to the wave equation) but is no longer justified when dealing with images, where no particular form of anisotropy can be postulated a priori. To the opposite, figuring out the precise form of anisotropy present in data is part of the issue. This argument also implies that one should not restrict analysis to tools that match one specific type of anisotropy, but rather that to tools embracing all of them simultaneously, in order to be able to detect that that suits data. From now on, two possible solutions for this problem will be focused on: \begin{itemize} \item One is supplied by {\em anisotropic Triebel bases}, see~\cite{triebel:2004}, that are constructed from the standard wavelet case through a multiresolution procedure, tailored to a specific anisotropy. The collection of these bases does not constitute a frame. However, for a fixed anisotropy, simple characterizations of anisotropic Besov spaces have been supplied within this system. Such characterizations can thus be used as a building step to construct a multifractal formalism \cite{benbraiek;benslimane:2011b}. This is further detailed in Section~\ref{s:multimulti}. Triebel bases provide a powerful tool to deduce results on anisotropic Besov spaces, for {\it a fixed anisotropy}. In particular, it enables to show that these spaces are isomorphic to the corresponding isotropic Besov spaces. Further, some results such as embeddings or profiles of Besov characteristics can be obtained, via the transference method proposed by H. Triebel. However, when it comes to understand the link between different forms of anisotropy - in term of function spaces by example - this tool remains of limited interest. Indeed, the knowledge of the expansion of a function in one basis gives a priori neither information about its expansion in an other basis nor about its belonging to all anisotropic Besov spaces. \item Another possible decomposition system is supplied by {\em hyperbolic wavelets}, introduced in various settings under different denominations (standard, rectangular or hyperbolic wavelet analysis) notably in image coding (see~\cite{Westerink_P_1989_phd_subband_ci}), numerical analysis (see~\cite{Beylkin_G_1991_j-comm-pure-appl-math_fast_wtna1}, \cite{Beylkin_G_1993_p-symp-appl-math_wavelets_fna}) and in~\cite{devore:konyagin:temlyakov:1998},\cite{hochmuth:2002a} for the purpose of approximation theory. They are simply defined as tensor products of 1D wavelets, yet allowing different dilations factors along different directions, as opposed to the classical discrete wavelet transform that relies on a single isotropic fixation factor. This key difference enables the study of anisotropy. Hyperbolic wavelet basis form a non--redundant system by construction, and contain all possible anisotropies. Hyperbolic wavelet bases have thus been used in statistics for the purpose of adaptive estimation of multidimensional curves. Notably, it has been proven in two seminal articles~\cite{neumann:vonsachs:1997} and~\cite{neumann:2000} that nonlinear thresholding of noisy hyperbolic wavelet coefficients leads to (near)--optimal minimax rates of convergence over a wide range of anisotropic smoothness classes. The reader is also referred to the recent work of F. Autin, G. Claeskens, J.M. Freyermuth \cite{autin:claeskens:freyermuth:2012} where this problem is considered from the maxiset point of view. Other interesting applications of hyperbolic analysis can also be founded in~\cite{ayache:leger:pontier:2002},\cite{ayache:2004},\cite{ayache:xiao:2005},\cite{ayache:roueff:xiao:2009a} and~\cite{ayache:roueff:xiao:2009b} where hyperbolic wavelet decompositions of Fractional Brownian Sheets and Linear Fractional Stable Sheets are given and are used to prove many sample paths properties of these random fields (smoothness properties, Hausdorff dimension of the graph). The key feature of hyperbolic wavelet bases is that they provide a common dictionary for anisotropic Besov spaces. This result is stated in Theorem~\ref{th:WCBesov} of Section~\ref{s:besov}: The critical exponent in anisotropic Besov spaces will be related to some $\ell^p$ norms of the hyperbolic wavelet coefficients. These mathematical results yield an efficient method for the detection of anisotropy, as detailed in a companion article, where numerical investigations are conducted, \cite{roux:clausel:vedel:jaffard:abry:2012} . \end{itemize} In the present article, it has been chosen to explore the possibilities supplied by the hyperbolic wavelet transform to investigate directional regularity, both in global (anisotropic Besov spaces) and local (directional pointwise regularity) forms. The underlying motivation is to develop a multifractal formalism relating these two notions (just as the standard multifractal formalism relates the usual Besov spaces with the notion of (anisotropic) H\"older pointwise smoothness, see~\cite{jaffard:2004} and references therein). It also aims at obtaining a numerically stable procedure that thus permits to extract the anisotropic features existing in natural images as well as information related to the size (fractional dimensions) of the corresponding geometrical sets. Before proceeding further, let us motivate the choice of hyperbolic wavelets against Triebel bases. For a fixed anisotropy, one can argue that Triebel bases display slightly better mathematical advantages: An exact characterization of anisotropic Besov spaces, as shown in~\cite{triebel:2004}, and a characterization of pointwise smoothness as sharp as in the isotropic case, as shown by H. Ben Braiek and M. Ben Slimane in~\cite{benbraiek:benslimane:2011a}. However, a first purpose of the present contribution is to show that these two important properties hold almost as well for hyperbolic wavelets: In Section~\ref{s:besov}, ``almost characterizations'' (i.e., necessary and sufficient conditions that differ by a logarithmic correction) of anisotropic Besov spaces are obtained. Furthermore, if one is not only interested in analysis, but also in simulation, this slight disadvantage (a logarithmic loss, which in applications can not be detected) is overcompensated by the advantage of using a basis instead of an overcomplete system. Indeed, generating a random field with prescribed regularity properties requires the use of a basis (using an overcomplete system cannot guarantee a priori that the simulated field with coefficients of specific sizes has the expected properties, since nontrivial linear combinations of the building blocks may vanish). A contrario, with the hyperbolic wavelet basis, one can easily provide toy examples with different multifractal spectra depending on the anisotropy. Our being jointly motivated by analysis and synthesis motivates the choice of a system that permits an interesting trade-off among directional wavelets, in terms of mathematical efficiency and numerical simplicity and robustness, both on the analysis and synthesis sides. The practical relevance of the mathematical tools introduced and studied here are assessed in a companion paper~\cite{roux:clausel:vedel:jaffard:abry:2012}. \\ Let us now further compare Triebel and hyperbolic wavelet bases in terms of pointwise directional smoothness. First, note that this notion has been the subject of few investigations so far: To our knowledge, the natural definition which allows for a wavelet characterization was first introduced by M. Ben Slimane in the 90s, see~\cite{benslimane:1998}, in order to investigate the multifractal properties of anisotropic selfsimilar functions. Partial results when using parabolic basis (i.e., curvelets and Hart--Smith transform) have been obtained by J. Sampo and S. Sumetkijakan see~\cite{lakhonchai:sampo:sumetkijakan:2012,sampo:sumetkijakan:2009,nualtong:sumetkijakan:2005} and references therein. A generalization and implications in terms of sizes of coefficients on directional wavelets (the so-called ``anisets'', which are a mixture of of the wavelet and Gabor transform, where the wavelets can be arbitrarily shrunk in certain directions) were also worked out in~\cite{jaffard:2010}. Finally, an ``almost '' characterization of pointwise directional regularity was recently obtained by H. Ben Braiek and M. Ben Slimane in~\cite{benbraiek;benslimane:2011a} on the Triebel basis coefficients, where the basis is picked so that its anisotropy parameter is fitted to the type of directional regularity considered. In Section~\ref{s:besov}, we will obtain a similar result, but relying on the coefficients of the hyperbolic wavelet basis, thus paving the way to the construction of a multifractal formalism. An important difference with~\cite{benbraiek;benslimane:2011a} is that, here, a single basis fits all anisotropies. Therefore, as in the case of Besov spaces, the advantage is that no a priori needs to be assumed on the particular considered anisotropy. This thus can be used as a way to detect the specific anisotropy which exists in data at hand, rather than assuming a priori its particular form beforehand. Note that other decomposition systems have also been used for the detection of local singularities, see for instance~\cite{donoho:1999,guo:labate:2011} where shearlets and wedgelets are used for the detection of discontinuities along smooth edges.\\ Let us now come back to the anisotropic self--similar fields considered in~\cite{roux:clausel:vedel:jaffard:abry:2012,clausel:vedel:2010}. Such exactly selfsimilar models are somewhat toy examples, and, though testing regularity indices on their realizations is an important validation step, their study could prove misleadingly simple (just as, in 1D, fractional Brownian motion is too simple a model to fit the richness of situations met in real-world data). Natural images are indeed likely to consist of patchworks of different kinds of deformed pieces and therefore, can be expected to exhibit more complex scale invariance properties, and only in an approximate way. A natural setting to describe such properties, where different kinds of singularities are mixed up, is supplied by multifractal analysis. The next step is therefore to combine both anisotropy and multifractal analyses. To this end, a new form of multifractal analysis is introduced, based on the hyperbolic wavelet coefficients, and relating the global and local characterizations of regularity. It allows to take into account both scale invariance properties and local anisotropic features of an image. Thus, it provides a new tool for image classification, seen as a refinement of texture classification based on the usual isotropic multifractal analysis, as proposed for instance in~\cite{abry:jaffard:wendt:2012},\cite{jaffard:2004}. Section~\ref{s:multimulti} is devoted to the introduction of this new framework: A new multifractal formalism, referred to as the {\it hyperbolic multifractal formalism}. It allows to relate local anisotropic regularity of the analyzed image to global quantities called hyperbolic structure functions as commonly done in multifractal analysis. Note that alternative multifractal analysis and multifractal formalism were introduced by H. Ben Braiek and M. Ben Slimane in~\cite{benbraiek;benslimane:2011b}, based on Triebel basis coefficients. In their approach, a particular anisotropy is picked, and the corresponding basis is used. As above, the main difference between our point of view and theirs is that we do not pick beforehand a particular anisotropy: Therefore, the approach proposed here does not rely on any a priori assumptions on data, and can thus be used when anisotropies of several types are simultaneously present in data. Finally, detailed proofs of all the results stated in Sections~\ref{s:besov} and~\ref{s:multimulti} are provided in Section~\ref{s:proofs}. \section{Anisotropic global regularity and hyperbolic wavelets}\label{s:besov} We first focus on the measure of anisotropic global regularity using a common analyzing dictionary: hyperbolic wavelet bases. Here, we start by providing the reader with a brief account of the corresponding functional spaces. Thereafter, we recall some well-known facts about hyperbolic wavelet analysis (cf. Section~\ref{s:WCBesov}). The main result of the present section consists of Theorem~\ref{th:WCBesov}, proven in Section~\ref{s:proofBesov}, which allows to determine the critical directional Besov indices of data by regressions on log-log plot of quantities based on hyperbolic wavelet coefficients (see Section~\ref{s:WCBesov} for a precise statement). \subsection{Anisotropic Besov spaces} Anisotropic Besov spaces generalize classical (isotropic) Besov spaces, and many results concerning isotropic spaces have been extended in this setting, see~\cite{bownik:ho:2005,bownik:2005} for a complete account on the results used in this section, and~\cite{bownik:2003,triebel:2006} for detailed overviews on anisotropic spaces. Note in particular that these spaces are invariant by smooth diffeomorphisms on each coordinate, an important requirement for image processing. Anisotropic Besov spaces verify (asymptotically in the limit of small scales) norm invariances with respect to anisotropic scaling, we, therefore, start by recalling this notion. Let $\alpha =(\alpha_1,\alpha_2)$ denote a fixed couple of parameters, with $\alpha_1, \, \alpha_2 \ge 0$ and $\alpha_1+\alpha_2 =2$. In the remainder, such couples will be referred to as {\em admisible anisotropies}. Such couples quantify the degree of anisotropy of the space ($\alpha_1= \alpha_2=1$ corresponding to the isotropic case). For any $t\ge 0$ and $\xi=(\xi_1,\xi_2) \in \mathbb{R}^2$, we define anisotropic scaling by $t^{\alpha} \xi = (t^{\alpha_1} \xi_1, t^{\alpha_2}\xi_2)$. Note that, in this definition and in the following, the coordinate axes are chosen as anisotropy directions. This particular choice can of course be modified by the introduction of an additional rotation (as envisaged e.g., in\cite{roux:clausel:vedel:jaffard:abry:2012}). Anisotropic Besov spaces may be introduced using an anisotropic Littlewood Paley analysis, which we now recall. Let $\varphi_0^{\alpha} \ge 0$ belong to the Schwartz class ${\mathcal{S}}(\mathbb{R}^2)$ and be such that $$ \varphi_0^{\alpha}(x) = 1 \quad if \quad \sup_{i=1,2} \vert \xi_{i} \vert \le 1\;, $$ and $$ \varphi_0^{\alpha}(x)=0 \quad if \quad \sup_{i=1,2} \vert 2^{-\alpha_{i}} \xi \vert \ge 1\;. $$ For $j \in \mathbb{N}$, we define $$ \varphi_j^{\alpha}(x) = \varphi_0^{\alpha}(2^{-j \alpha}\xi)-\varphi_0^{\alpha}(2^{-(j-1)\alpha}\xi)\;. $$ Then, $$ \sum_{j=0}^{+\infty} \varphi_j^{\alpha}\equiv 1\;, $$ and $(\varphi_j^{\alpha})_{j\geq 0}$ is called an {\it anisotropic resolution of the unity}. It satisfies $$ \mathrm{supp}\left(\varphi_0^{\alpha}\right) \subset R_1^{\alpha}, \quad \mathrm{supp}\left(\varphi_k^{\alpha}\right) \subset R_{j+1}^{\alpha} \setminus R_j^{\alpha}\;, $$ where $$ R_j^{\alpha} = \lbrace \xi=(\xi_1,\xi_2) \in \mathbb{R}^2;\, \sup_{i=1,2}\vert \xi_{\ell} \vert \le 2^{\alpha_i k}\rbrace\;. $$ For $f \in \mathcal{S}'(\mathbb{R}^2)$ let $$ \Delta^ {\alpha}_j f = {\mathcal{F}}^{-1} \left( \varphi^{\alpha}_j \widehat{f} \right)\;. $$ The sequence $(\Delta^ {\alpha}_j f )_{j \ge 0})$ is called an {\it anisotropic Littlewood--Paley analysis} of $f$. The anisotropic Besov spaces are then defined as follows (see~\cite{bownik:ho:2005,bownik:2005}). \begin{definition} The Besov space $B^{s,\alpha}_{p,q,|\log|^\beta}(\mathbb{R}^2)$, for $0<p \le +\infty$, $0<q\le + \infty$, $s,\beta \in \mathbb{R}$, is defined by $$ B^{s,\alpha}_{p,q,|\log|^\beta}(\mathbb{R}^2) = \lbrace f \in \mathcal{S}'(\mathbb{R}^2); \, \left( \sum_{j \ge 0} j^{-\beta q} 2^{jsq} \Vert \Delta^ {\alpha}_j f \Vert_p^q \right)^{1/q} <+\infty \rbrace\;. $$ This definition does not depend on the resolution of the chosen unity $\varphi_0^{\alpha}$ and the quantity $$ \Vert f \Vert_{B^{s,\alpha}_{p,q,|\log|^\beta}} = \left( \sum_{j \ge 0} j^{-\beta q} 2^{jsq} \Vert \Delta^ {\alpha}_jf \Vert_p^q \right)^{1/q}\;, $$ is a norm (resp., quasi-norm) on $B^{s,\alpha}_{p,q}(\mathbb{R}^2)$ for $1 \leq p, \, q \leq +\infty$ (resp., $0<p, \, q <1$). \end{definition} As in the isotropic case, anisotropic Besov spaces encompass a large class of classical anisotropic functional spaces (see~\cite{triebel:2006} for details). For example, when $p=q=2$ and $(\alpha_1,\alpha_2)\in\mathbb{Q}^2$ is an admissible anisotropy, let us consider $s>0$ such that $s/\alpha_1$ and $s/\alpha_2$ are both integers, then the anisotropic Sobolev space \[ H^{s,\alpha}(\mathbb{R}^2)=\{f \in L^2(\mathbb{R}^2)\mbox{ such that }\frac{\partial^{s/\alpha_1}f}{\partial x_1}\in L^2(\mathbb{R}^2)\mbox{ and }\frac{\partial^{s/\alpha_2}f}{\partial x_2}\in L^2(\mathbb{R}^2)\}\;, \] coincides with the Besov space $B_{2,2}^{s,\alpha}(\mathbb{R}^2)$. In the special case where $p=q=\infty$, the spaces $B^{s,\alpha}_{\infty,\infty}(\mathbb{R}^2)$ are called anisotropic H\"{o}lder spaces and are denoted $\mathcal{C}^{s,\alpha}_{|\log|^u}(\mathbb{R}^{2})$. These spaces also admit a finite difference characterization that we now recall (see also~\cite{triebel:2006} for details). Let $(e_1,e_2)$ denote the canonical basis of $\mathbb{R}^2$. For a function $f :\mathbb{R}^{2}\rightarrow \mathbb{R}$, $\ell\in\{1,2\}$ and $t\in\mathbb{R}$ one defines \[ \Delta^{1}_{t,\ell}f(x)=f(x+t e_\ell)-f(x)\;. \] The difference of order $M$, $M\geq 2$, of function $f$, along direction $e_\ell$, is then iteratively defined as \[ \Delta^M_{t,\ell}f(x)=\Delta_{t,\ell}\Delta^{M-1}_{t,\ell}f(x)\;. \] One then has: \begin{proposition} Let $\alpha=(\alpha_1,\alpha_2)\in (\mathbb{R}^+_*)^2$ such that $\alpha_1+\alpha_2=2$, $s>0$, $u\in\mathbb{R}$ and $f: \mathbb{R}^2\rightarrow \mathbb{R}$. The function $f$ belongs to the anisotropic H\"{o}lder space $\mathcal{C}^{s,\alpha}_{|\log|^u}(\mathbb{R}^{2})$ if \[ \|f\|_{L^{\infty}(\mathbb{R}^{2})}+\sum_{\ell=1}^2\sup_{t>0}\frac{\|\Delta^{M_\ell}_{t,\ell}f(x)\|_{L^{\infty}(\mathbb{R}^{2})}}{|t|^{s/\alpha_\ell}|\log(|t|)|^u}<+\infty\;, \] where for any $\ell\in\{1,2\}$, $M_\ell=[s/\alpha_\ell]+1$. \end{proposition} \subsection{Hyperbolic wavelet characterization of anisotropic Besov spaces}\label{s:WCBesov} We state our first main result which consists of an hyperbolic wavelet caracterization of anisotropic Besov spaces. We first need to recall the definition of the hyperbolic wavelet bases as tensorial products of two unidimensional wavelet bases (see~\cite{devore:konyagin:temlyakov:1998}) and second state Theorem~\ref{th:WCBesov}, further proven in Section~\ref{s:proofBesov}. \begin{definition} Let $\psi$ denote the unidimensional Meyer wavelet and $\varphi$ the associated scaling function. The hyperbolic wavelet basis is defined as the system $\{\psi_{j_1,j_2,k_1,k_2}, \, (j_1, j_2) \in (\mathbb{Z}^+\cup \{-1\})^2, \, (k_1, k_2) \in \mathbb{Z}^2\}$ where \begin{itemize} \item if $j_1,j_2\geq 0$, $$ \psi_{j_1,j_2,k_1,k_2}(x_1,x_2)= \psi(2^{j_1}x_1-k_1)\psi(2^{j_2}x_2-k_2)\;. $$ \item if $j_1=-1$ and $j_2\geq 0$ $$ \psi_{-1,j_2,k_1,k_2}(x_1,x_2)= \varphi(x_1-k_1)\psi(2^{j_2}x_2-k_2)\;. $$ \item if $j_1\ge 0$ and $j_2=-1$ $$ \psi_{j_1,-1,k_1,k_2}(x_1,x_2)= \psi(2^{j_1}x_1-k_1)\varphi(x_2-k_2)\;. $$ \item if $j_1=j_2=-1$ $$ \psi_{-1,-1,k_1,k_2}(x_1,x_2)= \varphi(x_1-k_1)\varphi(x_2-k_2)\;. $$ \end{itemize} For any $f\in\mathcal{S}'(\mathbb{R}^2)$, one then defines its hyperbolic wavelet coefficients as follows: \begin{eqnarray*} c_{j_1,j_2,k_1,k_2}&=&2^{j_1+j_2}<f,\psi_{j_1,j_2,k_1,k_2}>\mbox{ if }j_1,j_2\geq 0\;,\\ c_{j_1,-1,k_1,k_2}&=&2^{j_1}<f,\psi_{j_1,j_2,k_1,k_2}>\mbox{ if }j_1\geq 0\mbox{ and }j_2=-1\;,\\ c_{-1,j_2,k_1,k_2}&=&2^{j_2}<f,\psi_{j_1,j_2,k_1,k_2}>\mbox{ if }j_1=-1\mbox{ and }j_2\geq 0\;,\\ c_{-1,-1,k_1,k_2}&=&<f,\psi_{j_1,j_2,k_1,k_2}>\mbox{ if }j_1=j_2=-1\;. \end{eqnarray*} \end{definition} \begin{remark} We chose a $L^1$-normalization for the wavelet coefficients, known to be well-matching scale invariance. \end{remark} The main result of this section is an hyperbolic wavelet characterization of the spaces $B^{s,\alpha}_{p,q,|\log|^\beta}(\mathbb{R}^{2})$, up to a logarithmic correction. In the sequel, some notations are needed. For any $j=(j_1,j_2)\in (\mathbb{N}\cup\{-1\})^2$, let us define: \[ \|c_{j_1,j_2,\cdot,\cdot}\|_{\ell^p}=\left(\sum_{(k_1,k_2)\in\mathbb{Z}^2}|c_{j_1,j_2,k_1,k_2}|^p\right)^{1/p}\;. \] Let $\alpha=(\alpha_1,\alpha_2)$ be an admissible anisotropy, one also defines the following subsets of $\mathbb{N}^2$ \begin{eqnarray*} \Gamma_j^{(HL)}(\alpha)&=&\{(j_1,j_2)\in\mathbb{N}^2,\,[(j-1) \alpha_1 ]-1\leq j_1\leq [j\alpha_1]+1\mbox{ and }0\leq j_2\leq [(j-1) \alpha_2 ]-1\},\\ \Gamma_j^{(LH)}(\alpha)&=&\{(j_1,j_2)\in\mathbb{N}^2,\,0\leq j_1\leq [(j-1) \alpha_1 ]-1\mbox{ and }[(j-1) \alpha_2 ]-1\leq j_2\leq [j\alpha_2]+1\},\\ \Gamma_j^{(HH)}(\alpha)&=&\{(j_1,j_2)\in\mathbb{N}^2,\,[(j-1) \alpha_1 ]-1\leq j_1\leq [j\alpha_1]+1\mbox{ and }[(j-1) \alpha_2 ]-1\leq j_2\leq [j\alpha_2]+1\}\;, \end{eqnarray*} and \begin{equation}\label{e:gammaj} \Gamma_j(\alpha)=\Gamma_j^{(HL)}(\alpha)\cup \Gamma_j^{(LH)}(\alpha)\cup \Gamma_j^{(HH)}(\alpha)\;. \end{equation} Let us now state our hyperbolic wavelet characterization of anisotropic Besov spaces: \begin{theorem}\label{th:WCBesov} Let $\alpha=(\alpha_1,\alpha_2)$ be an admissible anisotropy, $(s,\beta)\in\mathbb{R}^2$ and $(p,q)\in (0,+\infty]^2$. Let $f\in \mathcal{S}'(\mathbb{R}^2)$. \begin{enumerate} \item Set $\beta(p,q)=\max(1/p-1,0)+\max(1-1/q,0)$. If \[ \left(\sum_{j\in\mathbb{N}_0}j^{q\beta(p,q)-\beta q}2^{jsq}\sum_{(j_1,j_2)\in \Gamma_j(\alpha)}2^{-\frac{(j_1+j_2)q}{p}}\;\|c_{j_1,j_2,\cdot,\cdot}\|_{\ell^p}^q\right)^{1/q}<+\infty\;, \] then $f \in B^{s,\alpha}_{p,q,|\log|^\beta}(\mathbb{R}^2)$ (with usual modifications when $q=\infty$). \item Conversely, \begin{enumerate} \item If $q<\infty$ and $f \in B^{s,\alpha}_{p,q,|\log|^\beta}(\mathbb{R}^2)$ then \[ \left(\sum_{j\in\mathbb{N}_0}j^{-\beta q-1}2^{jsq}\sum_{(j_1,j_2)\in \Gamma_j(\alpha)}2^{-\frac{(j_1+j_2)q}{p}}\;\|c_{j_1,j_2,\cdot,\cdot}\|_{\ell^p}^q\right)^{1/q}<+\infty\;. \] \item If $f \in B^{s,\alpha}_{p,\infty,|\log|^\beta}(\mathbb{R}^2)$ then \[ \max_{j\in\mathbb{N}_0}j^{-\beta }2^{js}\max_{(j_1,j_2)\in \Gamma_j(\alpha)}2^{-\frac{(j_1+j_2)}{p}}\;\|c_{j_1,j_2,\cdot,\cdot}\|_{\ell^p}<+\infty\;. \] \end{enumerate} \end{enumerate} \end{theorem} \begin{proof} Theorem~\ref{th:WCBesov} is proven in Section~\ref{s:proofBesov}. \end{proof} In particular, for $p=q=\infty$, the following ``almost'' characterization of anisotropic H\"{o}lder spaces by means of hyperbolic wavelets holds: \begin{proposition}\label{pro:WCGlobal} Let $\alpha=(\alpha_1,\alpha_2)$ an admissible anisotropy, $(s,\beta)\in\mathbb{R}^2$ and $f\in\mathcal{S}'(\mathbb{R}^2)$. \begin{enumerate}[(i)] \item If $f\in\mathcal{C}^{s,\alpha}(\mathbb{R}^{2})$ then there exists some $C>0$ such that for all $j\in \mathbb{N}\cup\{-1\}$ and any $(j_1,j_2)\in \Gamma_j(\alpha)$, \begin{equation}\label{e:WCGlobal1} \|c_{j_1,j_2,\cdot,\cdot}\|_{\ell^\infty}\leq C 2^{-js}\;. \end{equation} \item Conversely, assume that there exists some $C>0$ such that for all $j\in \mathbb{N}\cup\{-1\}$ and any $(j_1,j_2)\in \Gamma_j(\alpha)$ \begin{equation}\label{e:WCGlobal2} \|c_{j_1,j_2,\cdot,\cdot}\|_{\ell^\infty}\leq \frac{2^{- js}}{j}\;, \end{equation} then $f\in \mathcal{C}^{s,\alpha}(\mathbb{R}^{2})$. \end{enumerate} \end{proposition} In the special case where $p=q=2$, that is if we consider anisotropic Sobolev spaces, there is no logarithmic correction: \begin{theorem}\label{th:WCBesov2} Let $\alpha=(\alpha_1,\alpha_2)$ an admissible anisotropy, $s\in\mathbb{R}$. Let $f\in \mathcal{S}'(\mathbb{R}^2)$. The two following assertions are equivalent: \begin{enumerate}[(i)] \item $f \in H^{s,\alpha}(\mathbb{R}^2)=B^{s,\alpha}_{2,2}(\mathbb{R}^2)$. \item \[ \left(\sum_{j\in\mathbb{N}_0}2^{2js}\sum_{(j_1,j_2)\in\Gamma_j(\alpha)}2^{-(j_1+j_2)}\;\|c_{j_1,j_2,\cdot,\cdot}\|_{\ell^2}^2\right)^{1/2}<+\infty\;. \] \end{enumerate} \end{theorem} \begin{proof} Theorem~\ref{th:WCBesov2} is proven in Section~\ref{s:proofBesov}. \end{proof} \section{Hyperbolic multifractal analysis}\label{s:multimulti} We are now interested in the simultaneous analysis of local regularity properties and of anisotropic features of a function. To that end, we construct a new multifractal analysis, referred to as the hyperbolic multifractal analysis. Recall that in the isotropic case, the purpose of multifractal analysis is to provide information on the the pointwise singularities of functions. Multifractal functions are usually such that their local regularity strongly vary from point to point, so that it is not possible to estimate the regularity index (defined below) of a function at a given point. Instead, the relevant information consists of the ``sizes'' of the sets of points where the regularity index takes the same value. This ``size'' is mathematically formalized as the Hausdorff dimension. The function that associates the dimension of the set of points sharing the same regularity index with this index is referred to as the spectrum of singularities (or multifractal spectrum). The goal of a { \em multifractal formalism} is to provide a method that allows to measure the spectrum of singularities from quantities that can actually be computed on real-world data. We extend this approach to the anisotropic setting. Let us first recall that, in the case where the anisotropy of the analyzing space is fixed, this has already been achieved: See~\cite{benbraiek;benslimane:2011a} for anisotropic pointwise regularity analysis using Triebel bases and~\cite{benbraiek;benslimane:2011b} for the corresponding anisotropic multifractal formalism. Here, we generalize these two previous works, providing a multifractal analysis which does not rely on the a priori knowledge of the regularity and takes into account all possible anisotropies. Note that for a fixed anisotropy, both formalims coincide: Indeed they are derived from wavelet characterizations of the same functional spaces. Nevertheless, the formalism based on hyperbolic wavelet allows to deal simultaneously with all possible anisotropies, thus providing more useful algorithms for analyzing real-world data. In addition, the use of hyperbolic wavelet bases offers the possibility to define and synthesize deterministic and stochastic mathematical objects with prescribed anisotropic behavior. In Section~\ref{s:pointwise}, the different concepts related to pointwise regularity are first recalled. An hyperbolic wavelet criterion is then devised in Section~\ref{s:WLanis}. Our main result, Theorem~\ref{th:WL}, is stated in Section~\ref{s:WLanis} and proven in Section~\ref{s:proofs}. Hyperbolic wavelet analysis is defined in Section~\ref{s:hypmultiform} and the validity of the proposed multifractal formalism is investigated in Theorem~\ref{th:multform}. \subsection{Anisotropic pointwise regularity and hyperbolic wavelet analysis}\label{s:pointwise} \subsubsection{Definitions} Let us start by recalling the usual notion of pointwise regularity (cf.~\cite{jaffard:2004} for a complete review). \begin{definition} Let $f$ be in $L^\infty_{loc}(\mathbb{R}^2)$ and $s>0$. The function $f$ belongs to the space $\mathcal{C}^{s}_{|\log|^\beta}(x_0)$ if there exist some $C>0$, $\delta>0$ and $P_{x_0}$ a polynomial with degree less than $s$ such that \[ \mbox{if }|x-x_0|\leq \delta,\,|f(x)-P_{x_0}(x)|\leq C|x-x_0|^s\cdot |\log(|x-x_0|)|^\beta\;, \] where $|\cdot|$ is the usual Euclidean norm of $\mathbb{R}^{2}$. If $\beta=0$, the space $\mathcal{C}^{s}_{|\log|^0}(x_0)$ is simply denoted $\mathcal{C}^{s}(x_0)$. \end{definition} Anisotropic pointwise regularity is further defined as follows. Let $P$ denote a polynomial of the form: \[ P(t_1,t_2)=\sum_{(\beta_1,\beta_2)\in \mathbb{N}^2} a_{\beta_1,\beta_2} t_1^{\beta_1}t_2^{\beta_2}\;, \] and let $\alpha=(\alpha_1,\alpha_2)$ be an admissible anisotropy. The $\alpha$--homogeneous degree of the polynomial $P$ is defined as: \[ d_\alpha(P)=\sup\{\alpha_1 \beta_1+\alpha_2 \beta_2, a_{\beta_1,\beta_2}\neq 0\};. \] Finally, for any $t=(t_1,t_2)\in\mathbb{R}^2$, the $\alpha$--homogeneous norm reads: \[ |t|_\alpha=|t_1|^{1/\alpha_1}+|t_2|^{1/\alpha_2}\;. \] We can now define the spaces $\mathcal{C}^{s,\alpha}_{|\log|^\beta}(x_0)$. \begin{definition} Let $f\in L^\infty_{loc}(\mathbb{R}^2)$, $\alpha=(\alpha_1,\alpha_2)$ such that $\alpha_1+\alpha_2=2$, $|\cdot|_{\alpha}$, $s>0$ and $\beta\in\mathbb{R}$. The function $f$ belongs to $ \mathcal{C}^{s,\alpha}_{|\log|^\beta}(x_0)$ if there exists some $C>0$, $\delta>0$ and $P_{x_0}$ a polynomial with $\alpha$--homogeneous degree less than $s$ such that \[ \mbox{if }|x-x_0|_\alpha\leq \delta,\, |f(x)-P_{x_0}(x)|\leq C|x-x_0|_\alpha^s\cdot |\log(|x-x_0|_\alpha)|^\beta\;. \] If $\beta=0$, the space $\mathcal{C}^{s,\alpha}_{|\log|^0}(x_0)$ is simply denoted $\mathcal{C}^{s,\alpha}(x_0)$. \end{definition} The anisotropic pointwise exponent of a locally bounded function $f$ at $x_0$ can be thus be defined as: \begin{equation}\label{e:anisoexp} h_{f,\alpha}(x_0)=\sup\{s,\,f\in\mathcal{C}^{s,\alpha}(x_0)\}\;. \end{equation} The reader is referred to~\cite{benbraiek:benslimane:2011a},\cite{jaffard:2010} for more details about the material of this section. \subsubsection{An hyperbolic wavelet criterion}\label{s:WLanis} As in the usual anisotropic setting (see~\cite{jaffard:2004}), the anisotropic pointwise H\"older regularity of a function is closely related to the decay rate of decay of its wavelet leaders. The usual definition of wavelet leaders needs to be tuned to to the hyperbolic setting: For any $(j_1,j_2,k_1,k_2)$, let $\lambda(j_1,j_2,k_1,k_2)$ denote the hyperbolic dyadic cube: \begin{equation}\label{e:dyadiccube} \lambda=\lambda(j_1,j_2,k_1,k_2)=[\frac{k_1}{2^{j_1}},\frac{k_1+1}{2^{j_1}}[\times [\frac{k_2}{2^{j_2}},\frac{k_2+1}{2^{j_2}}[\;, \end{equation} and let $c_\lambda$ stand for $c_{j_1,j_2,k_1,k_2}$. The hyperbolic wavelet leaders $d_\lambda$, associated with the hyperbolic cube $\lambda$, can now be defined as: \[ d_{\lambda}=\sup_{\lambda'\subset \lambda}|c_{\lambda'}|\;. \] For any $x_0=(a,b) \in {\mathbb{R}}^2$, let \[ 3\lambda_{j_1,j_2}(x_0)=[\frac{[2^{j_1}a]-1}{2^{j_1}},\frac{[2^{j_1}a]+2}{2^{j_1}}[\times [\frac{[2^{j_2}b]-1}{2^{j_2}},\frac{[2^{j_2}b]+2}{2^{j_2}}[\;, \] (where $[\cdot]$ denotes the integer part) and \[ d_{j_1,j_2}(x_0)=\sup_{\lambda'\subset 3\lambda_{j_1,j_2}(x_0)}|c_{\lambda'}|\;. \] The hyperbolic wavelet leaders criterion for pointwise regularity can now be stated as: \begin{theorem}\label{th:WL} Let $s>0$ and $\alpha=(\alpha_1,\alpha_2)\in (\mathbb{R}^{*}_+)^2$ such that $\alpha_1+\alpha_2=2$. \begin{enumerate} \item Assume that $f\in \mathcal{C}^{s,\alpha}(x_0)$. There exists some $C>0$ such that for any $j_{1},j_2\in \mathbb{N}\cup \{-1\}$ one has \begin{equation}\label{e:WL} |d_{j_1,j_2}(x_0)|\leq C2^{-\max(\frac{j_1}{\alpha_1},\frac{j_2}{\alpha_2})s}\;. \end{equation} \item Conversely, assume that $f$ is uniformly H\"{o}lder and that~(\ref{e:WL}) holds, then $f\in \mathcal{C}^{s,\alpha}_{|\log|^2}(x_0)$. \end{enumerate} \end{theorem} Proofs are postponed to Section 4. \subsection{Anisotropic multifractal analysis} \subsubsection{Two notions of dimension} In multifractal analysis, two different notions of dimension are mainly used: the Hausdorff dimension and the packing dimension, whose definitions are now recalled. The Hausdorff dimension is defined through the Hausdorff measure (see~\cite{falconer:1990} for details). The best covering of a set $E\subset \mathbb{R}^d$ with sets subordinated to a diameter $\varepsilon$ can be estimated as follows, \[ \mathcal{H}_\varepsilon^\delta(E)=\inf\{\sum_{i=1}^\infty |E_i|^\delta,\,E\subset \bigcup_{i=1}^\infty E_i, |E_i|\leq \varepsilon\}. \] Clearly, $\mathcal{H}_\varepsilon^\delta$ is an outer measure. The Hausdorff measure is defined as the (possibly infinite or vanishing) limit $\mathcal{H}_\epsilon^\delta$ as $\varepsilon$ goes to $0$. The Hausdorff measure is decreasing as $\delta$ goes to infinity. Moreover, $\mathcal{H}^\delta(E)>0$ implies $\mathcal{H}^{\delta'}(E)=\infty$ if $\delta'<\delta$. The following definition is thus meaningful. \begin{definition} The Hausdorff dimension $\mathrm{dim}_H(E)$ of a set $E\subset \mathbb{R}^d$ is defined as follows, \[ \mathrm{dim}_H(E)=\sup\{\delta:\mathcal{H}^\delta(E)=\infty\}\;. \] \end{definition} With this definition, $\mathrm{dim}_H(\emptyset)=-\infty$. The box dimension (or Minkowski dimension) is simpler to define and to use than the Hausdorff dimension. \begin{definition} Let $E\subset \mathbb{R}^d$. If $\varepsilon>0$, let $N_\varepsilon(E)$ be the smallest number of sets of radius $\varepsilon$ required to cover $E$. The upper box dimension is \[ \overline{\mathrm{dim}}_B(E)=\limsup_{\varepsilon\to 0} \frac{\log N_\varepsilon(E)}{-\log \varepsilon}. \] The lower box dimension is \[ \underline{\mathrm{dim}}_B(E)=\liminf_{\varepsilon\to 0} \frac{\log N_\varepsilon(E)}{-\log \varepsilon}. \] If these two quantities are equal, they define the box dimension $\mathrm{dim}_B(E)$ of $E$. \end{definition} A significant limitation of box dimension is that a set and its closure have the same dimension. The packing dimension (introduced by Tricot, see~\cite{tricot:1991}) has better mathematical properties (e.g.,\ the packing dimension of a countable union of sets is the supremum of their dimensions). \begin{definition} The packing dimension $\mathrm{dim}_P(E)$ of a set $E\subset \mathbb{R}^d$ is defined by \[ \mathrm{dim}_P(E)=\inf\{\sup_i \{\overline{\mathrm{dim}}_B(E_i)\}:E\subset\bigcup_{i=1}^\infty E_i\}. \] \end{definition} The following inequality holds for any set $E\subset \mathbb{R}^d$, \[ \mathrm{dim}_H(E)\leq\mathrm{dim}_P(E). \] We now define the hyperbolic spectrum of singularities of a locally bounded function using the Hausdorff dimension. \begin{definition} Let $f$ be a locally bounded function and $\alpha=(\alpha_1,\alpha_2)\in (\mathbb{R}_{+}^*)^2$ such that $\alpha_1+\alpha_2=2$. The iso--anisotropic--H\"older set are defined as \[ E_f(H,\alpha)=\{x\in\mathbb{R}^2,h_{f,\alpha}(x)=H\} \] where the anisotropic pointwise H\"older $h_{f,\alpha}(x)$ has been defined in~(\ref{e:anisoexp}). The hyperbolic spectrum of singularities of $f$ is then defined as: \[ d: (\mathbb{R}^+ \cup \{\infty\})\times (0,2) \to \mathbb{R}^+ \cup \{-\infty\} \quad (H,a)\mapsto\mathrm{dim}_H(E_f(H,(a,2-a)))\;. \] \end{definition} \subsubsection{The hyperbolic wavelet leader multifractal formalism} \label{s:hypmultiform} It is not always possible to compute theoretically the spectrum of singularities of a given function. A multifractal formalism thus consists of a practical procedure that yields (a convex hull of) the function $d$, through the construction of structure functions and the use of the Legendre transform. In the classical case, these formalisms are variants of a seminal derivation, proposed by Parisi and Frisch in the context of the study of hydrodynamic turbulence~\cite{parisi:frisch:1985}. The hyperbolic wavelet leader multifractal formalism described below aimed at extending the procedure to where both anisotropy and singularities are studied jointly. Let us define hyperbolic partition functions of a locally bounded function as: \begin{equation}\label{eq:sf} S(j,p,\alpha)=2^{-2j}\sum_{(j_1,j_2)\in \Gamma_j(\alpha)} \sum_{(k_1,k_2)\in \mathbb{Z}^2}d_{j_1,j_2,k_1,k_2}^p\;, \end{equation} where $\Gamma_j(\alpha)$ has already been defined in Section~\ref{s:WCBesov} with Eq. (\ref{e:gammaj}). From the definition of an anisotropic scaling function (or scaling exponents) \begin{equation}\label{eq:omega} \omega_f(p,\alpha)=\liminf_{j\to\infty} \frac{\log S(j,p,\alpha)}{\log 2^{-j}}, \end{equation} let us further define the { \em Legendre hyperbolic spectrum}: \begin{equation}\label{eq:smf:leg} { \mathcal L}_f (H, \alpha) = \inf_{p\in\mathbb{R}^*}\{H p-\omega_f(p,(\alpha,2-\alpha))+2\}\;. \end{equation} Qualitatively, the Legendre hyperbolic spectrum and the hyperbolic spectrum of singularities $d_f(H,a)$ are expected to coincide, while the theorem below actually provides an upper bound relationship. \begin{theorem}\label{th:multform} Let $f$ a uniform H\"{o}lder function. The following inequality holds \begin{equation}\label{eq:majospectr} \forall (H,a)\in (\mathbb{R}^{*}_+)\times (0,2),\quad d_f(H,a)\leq { \mathcal L}_f (H, a). \end{equation} \end{theorem} \begin{definition} Let $f$ a uniform H\"{o}lder function, $H>0$ and $a\in (0,2)$. If Eq. (\ref{eq:majospectr}) simplifies into an equality, i.e., \[ \forall (H,a)\in (\mathbb{R}^{*}_+)\times (0,2),\quad d(H,a)= { \mathcal L}_f (H, \alpha) , \] then $f$ satisfies the hyperbolic multifractal formalism. \end{definition} From an applied perspective, Eqs. (\ref{eq:sf}) , (\ref{eq:omega}) and (\ref{eq:smf:leg}) constitute the core of the practical procedure enabling to compute the Legendre hyperbolic spectrum from the hyperbolic wavelet leaders computed on the data to be analyzed. Practical implementation show preliminary satisfactory results, notably, for isotropic function, it is clearly observed that the measured $ { \mathcal L}_f (H, \alpha) $ does not depend on $\alpha$. \section{Proofs}\label{s:proofs} \subsection{Proof of Theorem~\ref{th:WCBesov}}\label{s:proofBesov} \subsubsection{Hyperbolic Littlewood-Paley characterization of $B^{s,\alpha}_{p,q}(\mathbb{R}^2)$} \label{sec:p1} Let $\theta_0 \in \mathcal{S}(\mathbb{R},\mathbb{R}^+)$ be supported on $[-2,2]$ such that $\theta_0 =1$ on $[-1,1]$. For any $j \in \mathbb{N}$, let us define $$ \theta_j = \theta_0(2^j \cdot) - \theta_0(2^{j-1} \cdot)\;. $$ such that $\sum_{j \ge 0} \theta_j(\cdot) =1$ is a 1--D resolution of the unity. Observe that, for any $j \ge 1$, $\mathrm{supp}\left(\theta_j\right) \subset \{ 2^{j-1} \le \vert \xi \vert \le 2^{j+1} \}$. \begin{remark} In the following, the function $\theta_0$ can be chosen with an arbitrary compact support. It does not change the main results even if technical details of proofs and lemmas have to be adapted. It allows to chose the Fourier transform of a Meyer scaling function for $\theta_0$. \end{remark} \begin{definition} \, \begin{enumerate} \item For any $j, \, \ell \ge 0$, and any $\xi=(\xi_1,\xi_2) \in \mathbb{R}^2$ set $$ \phi_{j_1,j_2}(\xi_1, \xi_2)= \theta_{j_1}(\xi_1) \theta_{j_2}(\xi_2)\;. $$ For any $j_1, \, j_2 \ge 0$, the function $\phi_{j_1,j_2}$ belongs to $\mathcal{S}(\mathbb{R}^2)$ and is compactly supported on $\{2^{\ell_1} \le \vert \xi_1 \vert \le 2^{\ell_1+1} \} \times \{ 2^{\ell_2} \le \vert \xi_2 \vert \le 2^{\ell_2+1}]$. Further $\sum_{j_1 \ge 0} \sum_{j_2 \ge 0} \phi_{j_1,j_2} = 1$ and $(\phi_{j_1,j_2})_{(j_1,j_2)\in\mathbb{N}^2}$ is called an hyperbolic resolution of the unity. \item For $f \in \mathcal{S}'(\mathbb{R}^2)$ and $j_1, \, j_2 \ge 0$ set $$ \Delta_{j_1,j_2} f = {\mathcal{F}}^{-1} \left( \phi_{j_1,j_2} \widehat{f} \right)\;. $$ The sequence $((\Delta_{j_1,j_2} f)_{j_1,j_2 \ge 0})$ is called an hyperbolic Littlewood--Paley analysis of $f$. \end{enumerate} \end{definition} In the remainder of the section, we are given $\alpha=(\alpha_1,\alpha_2)$ a fixed anisotropy satisfying $\alpha_1+\alpha_2=2$ and $(\varphi^{\alpha}_j)_{j \ge 0}$ an anisotropic resolution of the unity. One then defines the following functions for any $j\geq 0$, \begin{equation}\label{def:gj} g_j^\alpha=\sum_{j_1,j_2\in \Gamma_j(\alpha)}\phi_{j_1,j_2}\;, \end{equation} where the sets $\Gamma_j(\alpha)$ have been defined in~(\ref{e:gammaj}). \begin{remark} Hyperbolic Littlewood--Paley analysis is used in the definition of spaces of mixed smoothness. We refer to~\cite{schmeisser:triebel:2004} for a study of these spaces and to~\cite{sickel:ullrich:2009} for their link with tensor products of Besov spaces and their hyperbolic wavelet characterizations. \end{remark} We now provide the reader with the following hyperbolic Littlewood--Payley characterization of anisotropic Besov spaces: \begin{theorem}\label{th:LPChar} Let $s\in\mathbb{R}$ and $(p,q)\in (0,+\infty]^2$. \begin{enumerate} \item \begin{enumerate} \item\label{condi} If $q<\infty$ and \begin{equation}\label{e:condi} \left(\sum_{j \ge 0} j^{q\max(1/p-1,0)+\max(q-1,0)}\cdot j^{-\beta q}2^{jsq} \sum_{(j_1,j_2)\in \Gamma_j(\alpha)} \left\Vert \Delta_{j_1,j_2}(f) \right\Vert_p^q \right)^{1/q} <+\infty\;, \end{equation} then $f \in B^{s,\alpha}_{p,q,|\log|^\beta}(\mathbb{R}^2)$. \item\label{condibis} If \begin{equation}\label{e:conditer} \max_{j \ge 0} \left(j^{\max(1/p-1,0)+1}\cdot j^{-\beta}2^{js} \max_{(j_1,j_2)\in \Gamma_j(\alpha)} \left\Vert \Delta_{j_1,j_2}(f) \right\Vert_p \right)<+\infty\;, \end{equation} then $f \in B^{s,\alpha}_{p,\infty,|\log|^\beta}(\mathbb{R}^2)$. \end{enumerate} \item \begin{enumerate} \item\label{condii} If $q<\infty$ and $f \in B^{s,\alpha}_{p,q,|\log|^\beta}(\mathbb{R}^2)$ then \[ \left(\sum_{j \ge 0} j^{-1}\cdot j^{-\beta q}2^{jsq} \sum_{(j_1,j_2)\in \Gamma_j(\alpha)} \left\Vert \Delta_{j_1,j_2}(f) \right\Vert_p^q \right)^{1/q} <+\infty\;. \] \item\label{condiibis} If $f \in B^{s,\alpha}_{p,\infty,|\log|^\beta}(\mathbb{R}^2)$ then \[ \max_{j \ge 0} \left( j^{-\beta}2^{js} \sum_{(j_1,j_2)\in \Gamma_j(\alpha)} \left\Vert \Delta_{j_1,j_2}(f) \right\Vert_p \right) <+\infty\;. \] \end{enumerate} \end{enumerate} \end{theorem} The proof of Theorem~\ref{th:LPChar} consists of several steps, beginning with \begin{lemma}\label{lem:1} \, \begin{enumerate} \item\label{intersi} For any $j \ge 0$ and any $(j_1,j_2)\not\in \Gamma_j(\alpha)$, one has \begin{equation}\label{e:intersi} \mathrm{supp}(\varphi^{\alpha}_j)\cap \mathrm{supp}(\phi_{j_1,j_2})=\emptyset\;. \end{equation} \item\label{intersii} For any $j \ge 0$ and any $\ell\not\in \{j-1,j,j+1\}$, one has \begin{equation}\label{e:intersii} \mathrm{supp}(g^{\alpha}_j)\cap \mathrm{supp}(\varphi_{\ell}^{\alpha})=\emptyset\;. \end{equation} \end{enumerate} \end{lemma} \begin{proof} Point~\ref{intersi} of the lemma is first proved, that is if ($\ell_1\ge L_{\max}^{(1)}+1$ and $\ell_2\ge L_{\max}^{(2)}+1$) or ($\ell_1\le L_{\min}^{(1)}-1$ and $\ell_2\le L_{\min}^{(2)}-1$), then $\mathrm{supp}(\varphi^{\alpha}_j)\cap \mathrm{supp}(\phi_{j_1,j_2})=\emptyset$. \\ Indeed, if $\xi\in \mathrm{supp}(\varphi^{\alpha}_j)$, then $\xi\in R_{j+1}^\alpha\setminus R_j^\alpha$. Hence, if $\ell_1\ge L_{\max}^{(1)}+1$ and $\ell_2\ge L_{\max}^{(2)}+1$, one has for $i=1,2$, $|2^{-\ell_i}\xi_i|\leq 2^{\alpha_i (j+1)-\ell_i}\leq 2^{\alpha_i-1} $, by assumptions on $\ell_1,\ell_2$. Since $\alpha_1$ or $\alpha_2$ necessarily belongs to $(0,1)$, one has $\xi\not\in \mathrm{supp}(\phi_{\ell_1,\ell_2})$. Hence, $\phi_{\ell_1,\ell_2}(\xi)=0$. The same approach leads to $\phi_{\ell_1,\ell_2}(\xi)=0$ if $\ell_1\le L_{\min}^{(1)}-1$ and $\ell_2\le L_{\min}^{(2)}-1$ and Point~(\ref{intersi}) of Lemma~\ref{lem:1} is obtained. Point~(\ref{intersii}) of Lemma~\ref{lem:1} can be obtained similarly. \end{proof} From Lemma~\ref{lem:1} an intermediate hyperbolic Littelwood Paley characterization of anisotropic Besov spaces is obtained. \begin{proposition}\label{pro:interm}Let $(p,q)\in (0,+\infty]^2$, $s,\beta\in\mathbb{R}$. the two following assertions are equivalent: \begin{enumerate} \item\label{condi} $f \in B^{s,\alpha}_{p,q,|\log|^\beta}(\mathbb{R}^2)$. \item\label{condii} \begin{equation}\label{e:condii} \left(\sum_{j \ge 0} j^{-\beta q}2^{jsq} \left\Vert \sum_{(j_1,j_2)\in \Gamma_j(\alpha)} [ \Delta_{j_1,j_2}(f)] \right\Vert_p^q \right)^{1/q}<+\infty\;. \end{equation} \end{enumerate} \end{proposition} \noindent{\it Proof of Proposition~\ref{pro:interm}.} Let us first show that $f \in B^{s,\alpha}_{p,q,|\log|^\beta}(\mathbb{R}^2)$ implies Inequality~(\ref{e:condii}) of Proposition~\ref{pro:interm}. To this end, Point~\ref{intersi} of Lemma~\ref{lem:1} is used to deduce that for any $j$ \[ \varphi_j^\alpha \widehat{f}=\varphi_j^\alpha\left(\sum_{(j_1,j_2)\in\mathbb{N}^2}\phi_{j_1,j_2}\right)\widehat{f}=\varphi_j^\alpha\left(\sum_{(j_1,j_2)\in\Gamma_j(\alpha)}\phi_{j_1,j_2}\right)\widehat{f} =\varphi_j^\alpha\left(g_j^\alpha \widehat{f}\right)\;, \] where $g_j^\alpha$ is defined by Equation~(\ref{def:gj}). Observe now that replacing the usual dilation with an anisotropic one gives an anisotropic version of Equation~(13) in Section 1.5.2 in~\cite{triebel:1978}. More precisely assume that we are given $p\in (0,+\infty]$, $b>0$ and $M\in \mathcal{S}(\mathbb{R}^{2})$. There exists some $C>0$ not depending on $b$ nor $M$ such that for any $h\in L^p(\mathbb{R}^{2})$ such that $\mathrm{supp}(\widehat{h})\subset \{\xi\in \mathbb{R}^{2},\,\sup_i|\xi_i|\leq b^{\alpha_i}\}$, one has \begin{equation}\label{e:convmod} \|\mathcal{F}^{-1}\left(M\mathcal{F}h\right)\|_{L^p(\mathbb{R}^2)}\leq C\|M(b^\alpha\cdot)\|_{H_2^s(\mathbb{R}^2)}\|h\|_{L^p(\mathbb{R}^2)} \end{equation} where $H^s_2$ is the usual Bessel potential space and $s>2(1/\min(p,1)-1/2)$. Set now $b=2^j$, $M=\varphi_j^\alpha$ and $\widehat{h}=g_j^\alpha \widehat{f}$. Since $\varphi_j^\alpha(2^{j\alpha}\cdot)=\varphi_1^\alpha$, there exists some $C>0$ not depending on $j$ such that for any $p\in (0,+\infty]$ and any $f\in L^p(\mathbb{R}^{2})$ \[ \|\Delta_j^\alpha f\|_{L^p}\leq C\|\sum_{(j_1,j_2)\in\Gamma_j(\alpha)}\Delta_{j_1,j_2}f\|_{L^p} =C\|(\mathcal{F}^{-1}g_j^\alpha)*f\|_{L^p}\;. \] Then \begin{equation}\label{ineq:1} \|f\|_{B^{s,\alpha}_{p,q,|\log|^\beta}}=\left(\sum_{j\geq 0}j^{-\beta q}2^{jsq}\|\Delta_j^\alpha f\|_{L^p}^q\right)^{1/q} \leq C \left(\sum_{j\geq 0}j^{-\beta q} 2^{jsq}\|(\mathcal{F}^{-1}g_j^\alpha)*f\|_{L^p}^q\right)^{1/q}\;, \end{equation} which gives Point~\ref{condi} of Proposition~\ref{pro:interm}. Let us now prove that if Equation~(\ref{e:condii}) of Proposition~\ref{pro:interm} holds then $f\in B^{s,\alpha}_{p,q,|\log|^\beta}(\mathbb{R}^2)$. Point~\ref{intersii} of Lemma~\ref{lem:1} gives for any $j\geq 0$ \[ g_{j}^\alpha \widehat{f}=g_{j}^\alpha \left(\varphi_{j-1}^\alpha+\varphi_{j}^\alpha+\varphi_{j+1}^\alpha\right)\widehat{f}\;. \] Hence, Inequality~(\ref{e:convmod}) applied with $b=2^j$, $M=g_j^\alpha$ and $\widehat{h}=(\varphi_{j-1}^\alpha+\varphi_j^\alpha+\varphi_{j+1}^\alpha) \widehat{f}$ gives the existence of some $C>0$ not depending on $j$ nor $f$ such that for any $p\in (0,+\infty]$ \[ \|\left(\mathcal{F}^{-1}g_{j}^\alpha\right)*f\|_{L^p}\leq c\|g_j^\alpha(2^{j\alpha}\cdot)\|_{H_2^s}\|(\mathcal{F}^{-1}\varphi_{j-1}^\alpha+\mathcal{F}^{-1}\varphi_{j}^\alpha+\mathcal{F}^{-1}\varphi_{j+1}^\alpha)*f\|_{L^p} \] Since $\|\cdot\|_{L^p}$ is either a norm or a quasi--norm (according to the value of $p$), there exists some $C>0$ such that \begin{eqnarray*} \|\left(\mathcal{F}^{-1}g_{j}^\alpha\right)*f\|_{L^p}\leq & C & \|g_j^\alpha(2^{j\alpha}\cdot)\|_{H_2^s} \\ & & \left(\|(\mathcal{F}^{-1}\varphi_{j-1}^\alpha)*f\|_{L^p}+ \|(\mathcal{F}^{-1}\varphi_{j}^\alpha)*f\|_{L^p}+\|(\mathcal{F}^{-1}\varphi_{j+1}^\alpha)*f\|_{L^p}\right)\;. \end{eqnarray*} Let us first bound $\|g_j^\alpha(2^{j\alpha}\cdot)\|_{H_2^s}$. To this end, observe that \begin{eqnarray*} \mathcal{F}[g_j^\alpha(2^{j\alpha}\cdot)](\xi) & = & 2^{-j(\alpha_1+\alpha_2)}\widehat{g_j}(2^{-j\alpha}\xi)=2^{-2j}\widehat{g_j}(2^{-j\alpha}\xi) \\ & =&\sum_{(j_1,j_2)\in \Gamma_j(\alpha)}2^{j_1+j_2-2j}\widehat{\theta}_{1}(2^{j_1-j\alpha_1}\,\xi_1)\widehat{\theta}_{1}(2^{j_2-\alpha_2 j}\,\xi_2)\;. \end{eqnarray*} Hence \begin{eqnarray*} \|g_j^\alpha(2^{j\alpha}\cdot)\|_{H_2^s}^2&=&\int_{\mathbb{R}^2}(1+|\xi|^2)^s\left[\sum_{(j_1,j_2)\in \Gamma_j(\alpha)}2^{j_1+j_2-2j}\widehat{\theta}_{1}(2^{j_1-j\alpha_1}\,\xi_1)\widehat{\theta}_{1}(2^{j_2-\alpha_2 j}\,\xi_2)\right]^2\mathrm{d} \xi\\ &\leq&\int_{\mathbb{R}^2}(1+|\xi|^2)^s\left[\sum_{(j_1,j_2)\in \Gamma_j(\alpha)}2^{j_1+j_2-2j}|\widehat{\theta}_{1}(2^{j_1-j\alpha_1}\,\xi_1)||\widehat{\theta}_{1}(2^{j_2-\alpha_2 j}\,\xi_2)|\right]^2\mathrm{d} \xi\;. \end{eqnarray*} Since $\theta_1\in \mathcal{S}(\mathbb{R})$, for any $M>1$ there exists some $C>0$ such that \[ |\widehat{\theta}_1(\zeta)|\leq \frac{C_M}{(1+|\zeta|)^{2M}}\;. \] Finally \begin{eqnarray*} \|g_j^\alpha(2^{j\alpha}\cdot)\|_{H_2^s}^2 &\leq & C_M\int_{\mathbb{R}^2}(1+|\xi|^2)^s \\ & & \hspace{1.5cm} \times \left[\sum_{(j_1,j_2)\in \Gamma_j(\alpha)} \frac{2^{j_1+j_2-2j}}{(1+|2^{j_1-j\alpha_1}\,\xi_1|)^{2M} \cdot (1+|2^{j_2-j\alpha_2}\,\xi_2|)^{2M}}\right]^2\mathrm{d} \xi\\ &\leq& C_M\int_{\mathbb{R}^2}(1+|\xi|^2)^s \\ & & \hspace{1.5cm} \times \left[\sum_{(j_1,j_2)\in \Gamma_j(\alpha)} \frac{1}{(2^{j\alpha_1-j_1}+|\xi_1|)^{2M} \cdot (2^{j\alpha_2-j_2}+|\xi_2|)^{2M}}\right]^2\mathrm{d} \xi\;. \end{eqnarray*} By the inequality \[ (a+b)^2\geq a\max(b,1)\;, \] valid for any $a>1$, $b>0$ and applied successively with $a=2^{j\alpha_1-j_1}$ and $b=|\xi_1|$, $a=2^{j\alpha_2-j_2}$ and $b=|\xi_2|$, it comes \begin{eqnarray*} \|g_j^\alpha(2^{j\alpha}\cdot)\|_{H_2^s}^2&\leq & C_M \int_{\mathbb{R}^2}(1+|\xi|^2)^s \\ & & \times \left[\sum_{(j_1,j_2)\in \Gamma_j(\alpha)}2^{(j_1-j\alpha_1)M}2^{(j_2-j\alpha_2)M} \times\frac{1}{\max(1,|\xi_1|)^{M}\max(1,|\xi_2|)^{M}}\right]^2\mathrm{d} \xi\;. \end{eqnarray*} With a $M$ sufficiently large it follows that \[ \sup_j\left(\|g_j^\alpha(2^{j\alpha}\cdot)\|_{H_2^s}\right)<+\infty\;. \] Going back to an upper bound of $\|\left(\mathcal{F}^{-1}g_{j}^\alpha\right)*f\|_{L^p}$, there exists some $C>0$ such that \[ \|\left(\mathcal{F}^{-1}g_{j}^\alpha\right)*f\|_{L^p}\leq Cj\left(\|(\mathcal{F}^{-1}\varphi_{j-1}^\alpha)*f\|_{L^p}+ \|(\mathcal{F}^{-1}\varphi_{j}^\alpha)*f\|_{L^p}+\|(\mathcal{F}^{-1}\varphi_{j+1}^\alpha)*f\|_{L^p}\right)\; \] and \begin{equation}\label{ineq:2} \left(\sum_{j\geq 0}j^{-\beta q}2^{jsq}\|(\mathcal{F}^{-1}g_j^\alpha)*f\|_{L^p}^q\right)^{1/q}\leq C\|f\|_{B^{s,\alpha}_{p,q,|\log|^{\beta}}}=\left(\sum_{j\geq 0}j^{-\beta q} 2^{jsq}\|\Delta_j^\alpha f\|_{L^p}^q\right)^{1/q}\;, \end{equation} the last shows that if (\ref{e:condii}) holds then $f\in B^{s,\alpha}_{p,q,|\log|^{\beta}}(\mathbb{R}^2)$.\\ \noindent{\it Proof of Theorem~\ref{th:LPChar}.} Let us first recall that: \begin{itemize} \item For any $p\in (0,+\infty]$, $n\in\mathbb{N}$, and $(f_1,\cdots,f_n)\in L^{p}(\mathbb{R}^2)^n$ \begin{equation}\label{ineq:cvxity1} \|f_1+\cdots+f_n\|_{L^p}\leq n^{\max(1/p-1,0)}\left(\|f_1\|+\cdots+\|f_n\|\right) \end{equation} \item For any $q\in (0,+\infty)$, $n\in\mathbb{N}$, and $(a_1,\cdots,a_n)\in (\mathbb{R}_+)^n$ \begin{equation}\label{ineq:cvxity2} (a_1+\cdots+a_n)^q\leq n^{\max(q-1,0)}\left(a_1^q+\cdots+a_n^q\right)\;. \end{equation} \end{itemize} Let us now prove the first point of the theorem in the case where $q\neq\infty$. For this, assume that~(\ref{e:condi}) holds and let us prove that $f\in B^{s,\alpha}_{p,q,|\log|^{\beta}}(\mathbb{R}^2)$. By Inequalities~(\ref{ineq:cvxity1}), (\ref{ineq:cvxity2}) and the fact that $\mathrm{Card}(\Gamma_j(\alpha))\leq Cj$ there exists $C>0$ such that \[ \left\|\sum_{(j_1,j_2)\in \Gamma_j(\alpha)}\Delta_{j_1,j_2}f\right\|_{L^p}^q\leq C j^{q\max(1/p-1,0)+\max(q-1,0)}\sum_{(j_1,j_2)\in \Gamma_j(\alpha)}\|\Delta_{j_1,j_2}f\|_{L^p}^q\;. \] Hence, \begin{eqnarray*} &&\left(\sum_{j\geq 0}j^{-\beta q}2^{jsq}\left\|\sum_{(j_1,j_2)\in \Gamma_j(\alpha)}\Delta_{j_1,j_2}f\right\|_{L^p}^q\right)^{1/q}\\ &\leq& C \left(\sum_{j\geq 0}j^{q\max(1/p-1,0)+\max(q-1,0)}\cdot j^{-\beta q}2^{jsq}\sum_{(j_1,j_2)\in \Gamma_j(\alpha)}\|\Delta_{j_1,j_2}f\|_{L^p}^q\right)^{1/q}\;. \end{eqnarray*} It proves that if~(\ref{e:condi}) holds, one has \[ \left(\sum_{j\geq 0}j^{-\beta q} 2^{jsq}\|\sum_{(j_1,j_2)\in \Gamma_j(\alpha)}\Delta_{j_1,j_2}f\|_{L^p}^q\right)^{1/q}<\infty\;. \] Finally, by Point~(1) of Proposition~\ref{pro:interm}, it comes that $f\in B^{s,\alpha}_{p,q,|\log|^{\beta}}(\mathbb{R}^2)$. \noindent We now deal with the case $q=\infty$. In this case, we have \begin{eqnarray*} \max_{j\geq 0}j^{-\beta }2^{js}\left\|\sum_{(j_1,j_2)\in \Gamma_j(\alpha)}\Delta_{j_1,j_2}f\right\|_{L^p}\leq C\max_{j\geq 0}j^{\max(1/p-1,0)}j^{-\beta }2^{js}\sum_{(j_1,j_2)\in \Gamma_j(\alpha)}\|\Delta_{j_1,j_2}f\|_{L^p}\;. \end{eqnarray*} Hence if~(\ref{e:condii}) holds, $f\in B^{s,\alpha}_{p,\infty,|\log|^{\beta}}(\mathbb{R}^2)$. To prove the converse assertion, let us assume $f\in B^{s,\alpha}_{p,q,|\log|^\beta}(\mathbb{R}^2)$. Observe that for any $j\geq 0$ and any $(j_1,j_2)\in\Gamma_j(\alpha)$, one has \[ \phi_{j_1,j_2}\widehat{f}= \phi_{j_1,j_2}\left(g_{j-1}^\alpha+g_{j}^\alpha+g_{j+1}^\alpha\right)\widehat{f}\;. \] Remark that $\phi_{j_1,j_2}(2^{j\alpha}\cdot)$ is bounded in $H^s_2(\mathbb{R}^2)$ independently of $(j_1,j_2)\in \Gamma_j(\alpha)$. Hence, by ~(\ref{e:convmod}), there exists $C>0$ not depending on $j$ nor $f$ such that for any $(j_1,j_2)\in \Gamma_j(\alpha)$ \begin{eqnarray*} \|(\mathcal{F}^{-1}\phi_{j_1,j_2})*f\|_{L^p}&\leq& C\left(\|(\mathcal{F}^{-1}g_{j-1}^\alpha)*f\|_{L^p}+\|(\mathcal{F}^{-1}g_{j}^\alpha)*f\|_{L^p} +\|(\mathcal{F}^{-1}g_{j+1}^\alpha)*f\|_{L^p}\right)\;. \end{eqnarray*} Again, two cases have to be distinguished according whether $q\neq \infty$ or $q=\infty$. Let us consider the case $q<\infty$. Observing that $\mathrm{Card}(\Gamma_j(\alpha))\leq Cj$, we deduce that \[ \sum_{(j_1,j_2)\in\Gamma_j(\alpha)}\|(\mathcal{F}^{-1}\phi_{j_1,j_2})*f\|_{L^p}^q\leq Cj \sum_{l=j-1}^{j+1} \|(\mathcal{F}^{-1}g_{l}^\alpha)*f\|_{L^p}^q\;. \] So \begin{equation}\label{e:ineqa} \sum_j j^{-1}j^{-\beta q}2^{js q}\sum_{(j_1,j_2)\in\Gamma_j(\alpha)}\|(\mathcal{F}^{-1}\phi_{j_1,j_2})*f\|_{L^p}^q\leq \sum_j j\cdot j^{-1}j^{-\beta q}2^{js q}\|(\mathcal{F}^{-1}g_{j}^\alpha)*f\|_{L^p}^q\;. \end{equation} Since in addition the function $f$ is assumed to belong to $B^{s,\alpha}_{p,q,|\log|^\beta}(\mathbb{R}^2)$, one has \[ \sum_j j^{-\beta q}2^{js q}\|(\mathcal{F}^{-1}g_{j}^\alpha)*f\|_{L^p}^q=\sum_j j\cdot j^{-1}j^{-\beta q}2^{js q}\|(\mathcal{F}^{-1}g_{j}^\alpha)*f\|_{L^p}^q<\infty\;, \] which directly yields the required inequality using~(\ref{e:ineqa}). In the case $q=\infty$, we have \begin{eqnarray*} \max_{(j_1,j_2)\in \Gamma_j(\alpha)}\|(\mathcal{F}^{-1}\phi_{j_1,j_2})*f\|_{L^p}^q&\leq& C\;\max_{\ell=j-1,j,j+1}\|(\mathcal{F}^{-1}g_{\ell}^\alpha)*f\|_{L^p}\;. \end{eqnarray*} which leads for some $C>0$ to \[ \max_{j\geq 0}\left(j^{-\beta }2^{js}\max_{(j_1,j_2)\in \Gamma_j(\alpha)}\|(\mathcal{F}^{-1}\phi_{j_1,j_2})*f\|_{L^p}\right)\leq C\max_{j\geq 0}\left(j^{-\beta }2^{js}\|(\mathcal{F}^{-1}g_{j}^\alpha)*f\|_{L^p}\right)\;, \] that is \begin{equation}\label{ineq:pointiith} \max_{j_1,j_2\geq 0}\left(\max(\frac{j_1}{\alpha_1},\frac{j_2}{\alpha_2})\right)^{-\beta}2^{\max(\frac{j_1}{\alpha_1},\frac{j_2}{\alpha_2})s}\|(\mathcal{F}^{-1}\phi_{j_1,j_2})*f\|_{L^p}\leq C\max_{j\geq 0}j^{-\beta }2^{js}\|(\mathcal{F}^{-1}g_{j}^\alpha)*f\|_{L^p}\;. \end{equation} Since in addition $f$ is assumed to belong to $B^{s,\alpha}_{p,\infty,|\log|^\beta}(\mathbb{R}^2)$, it comes \[ \max j^{-\beta }2^{js}\|(\mathcal{F}^{-1}g_{j}^\alpha)*f\|_{L^p}<\infty\;. \] Finally, the required conclusion is obtained by an approach similar to the one used for the previous case. \subsubsection{The special case $p=2$}\label{s:thWC2} In that case, an {\bf exact} hyperbolic Littlewood--Paley characterization of anisotropic Besov spaces is provided: \begin{proposition}\label{pro:caracLP2} Let $f\in\mathcal{S}'(\mathbb{R}^2)$, $s>0$ and $q\in (0,+\infty]$. The two following assertions are equivalent \begin{enumerate}[(i)] \item $f \in B^{s,\alpha}_{2,q,|\log|^\beta}(\mathbb{R}^2)$. \item $\sum\limits_{j \ge 0}j^{-\beta q} 2^{jsq}\left(\sum\limits_{(j_1,j_2)\in \Gamma_j(\alpha)} \Vert \phi_{j_1,j_2} \widehat{f} \Vert^2_{L^2} \right)^{\frac{q}{2}}=\sum\limits_{j \ge 0} j^{-\beta q} 2^{jsq} \left( \sum\limits_{(j_1,j_2)\in \Gamma_j(\alpha)}\Vert \Delta_{j_1,j_2}(f) \Vert^2_{L^2} \right)^{\frac{q}{2}} <+\infty$. \end{enumerate} \end{proposition} In particular the following exact hyperbolic Littlewood--Paley characterization of anisotropic Sobolev spaces can be stated: \begin{theorem}\label{pro:caracLP2} Let $f\in\mathcal{S}'(\mathbb{R}^2)$, $s>0$ and $q\in (0,+\infty]$. The two following assertions are equivalent \begin{enumerate}[(i)] \item $f \in H^{s,\alpha}_{|\log|^\beta}(\mathbb{R}^2)=B^{s,\alpha}_{2,2}(\mathbb{R}^2)$. \item $\sum\limits_{j \ge 0} j^{-2\beta}2^{2js} \sum\limits_{(j_1,j_2)\in \Gamma_j(\alpha)} \Vert \phi_{j_1,j_2} \widehat{f} \Vert^2_{L^2} =\sum\limits_{j \ge 0} j^{-2\beta} 2^{2js} \sum\limits_{(j_1,j_2)\in \Gamma_j(\alpha)}\Vert \Delta_{j_1,j_2}(f) \Vert^2_{L^2}<+\infty$. \end{enumerate} \end{theorem} To prove Proposition~\ref{pro:caracLP2}, let us first precise the relation between anisotropic and hyperbolic resolutions of the unity. \begin{lemma}\label{lem:1} For any $j \geq 0$, the following inequality holds on $\mathbb{R}^2$ \begin{equation}\label{e:compLPHyp} \varphi^{\alpha}_j \le g_j=\sum_{(j_1,j_2)\in\Gamma_j(\alpha)}\phi_{j_1,j_2} \le \varphi^{\alpha}_{j-1}+\varphi^{\alpha}_j + \varphi^{\alpha}_{j+1}\;. \end{equation} \end{lemma} \begin{proof} Let us first observe that \[ \varphi^{\alpha}_j \le \sum_{(j_1,j_2)\in \mathbb{N}^2} \phi_{j_1,j_2}\;. \] To get the left hand side of inequality~(\ref{e:compLPHyp}), we just have to prove that if $\xi=(\xi_1,\xi_2)\in \mathrm{supp}(\varphi^{\alpha}_j)$, one has $\phi_{j_1,j_2}(\xi)=0$ if $(j_1,j_2)\not\in \Gamma_j$, that is if ($\ell_1\ge L_{\max}^{(1)}+1$ and $\ell_2\ge L_{\max}^{(2)}+1$) or ($\ell_1\le L_{\min}^{(1)}-1$ and $\ell_2\le L_{\min}^{(2)}-1$). \\ Indeed, if $\xi\in \mathrm{supp}(\varphi^{\alpha}_j)$, then $\xi\in R_{j+1}^\alpha\setminus R_j^\alpha$. Hence, if $\ell_1\ge L_{\max}^{(1)}+1$ and $\ell_2\ge L_{\max}^{(2)}+1$, one has for $i=1,2$, $|2^{-\ell_i}\xi_i|\leq 2^{\alpha_i (j+1)-\ell_i}\leq 2^{\alpha_i-1} $, by assumptions on $\ell_1,\ell_2$. Since $\alpha_1$ or $\alpha_2$ necessarily belongs to $(0,1)$, one has $\xi\not\in \mathrm{supp}(\phi_{\ell_1,\ell_2})$. Hence, $\phi_{\ell_1,\ell_2}(\xi)=0$. The same approach leads to $\phi_{\ell_1,\ell_2}(\xi)=0$ if $\ell_1\le L_{\min}^{(1)}-1$ and $\ell_2\le L_{\min}^{(2)}-1$. The left hand side of inequality~(\ref{e:compLPHyp}) is obtained. Let us now prove that the right hand side of inequality~(\ref{e:compLPHyp}) holds. It comes from the obvious equality \[ \sum_{(j_1,j_2)\in \Gamma_j} \phi_{j_1,j_2}\leq \sum_{j\ge 0}\varphi^{\alpha}_j\equiv 1 \;. \] and if $\xi\in \mathrm{supp}(\phi_{j_1,j_2})$ for some $(j_1,j_2)\in\Gamma_j(\alpha)$ then for any $\ell\in \{j-1,j,j+1\}$, $\varphi^{\alpha}_{\ell}(\xi)=0$.\\ \end{proof} Before proving Proposition~\ref{pro:caracLP2}, let us give some background about quasi--orthogonal systems. \begin{definition} Let a Hilbert space $H$ and $\langle\cdot,\cdot\rangle$ the associated scalar product. A system $\{f_k, k \in \mathbb{Z}\}$ of $H$ is said to be quasi-orthogonal if there exists some $\ell \in \mathbb{N}$ such that \begin{equation}\label{e:quasiorth} \forall (k,k') \in \mathbb{Z}^2, \, \left(\vert k'-k \vert \geq \ell \Rightarrow\quad \langle f_k, f_{k'} \rangle = 0\right)\;. \end{equation} \end{definition} \begin{lemma}\label{lem:qo} Let $H$ be a Hilbert space and $\|\cdot\|$,$\langle\cdot,\cdot\rangle$ the associated norm and scalar product. Let $\lbrace f_m, \, m \in \mathbb{Z}\}$ a quasi--orthogonal system of $H$ and let $\ell \in \mathbb{N}$ satisfying (\ref{e:quasiorth}). Then \begin{equation} \left\Vert \sum_{m \in \mathbb{Z}} f_m \right\Vert ^2 \leq (2\ell+1) \sum_{m \in \mathbb{Z}} \Vert f_m \Vert^2\;. \end{equation} \end{lemma} \begin{proof} Observe that for any $m' \in \mathbb{Z}$, $\langle f_m, f_{m'} \rangle =0$ except if $m'-\ell \le m \le m'+\ell$. Hence \begin{eqnarray*} \left\Vert \sum_{m \in \mathbb{Z}} f_m \right\Vert ^2 & \le & \sum_{m' \in \mathbb{Z}} \sum_{m \in \mathbb{Z}} \vert \langle f_m, f_{m'} \rangle \vert \\ & \le & \sum_{m' \in \mathbb{Z}} \sum_{m=m' -\ell}^{m'+\ell} \Vert f_m \Vert \Vert f_{m'} \Vert \\ & \le & \sum_{m' \in \mathbb{Z}} \sqrt{2\ell+1} \left( \sum_{m =m'-\ell}^{m'+\ell} \Vert f_m \Vert^2 \right)^{\frac{1}{2}} \Vert f{_m'} \Vert \\ & \le & \sqrt{2\ell+1} \left( \sum_{m' \in \mathbb{Z}} \left( \sum_{m=m'-\ell}^{m'+\ell} \Vert f_m \Vert^2 \right) \right)^{\frac{1}{2}}. \left( \sum_{m' \in \mathbb{Z}} \Vert f_{m'} \Vert^2 \right)^{\frac{1}{2}}\\ & \le & (2\ell+1)\sum_{m' \in \mathbb{Z}} \Vert f_{m'} \Vert^2\;. \end{eqnarray*} \end{proof} {\bf Proof of Proposition~\ref{pro:caracLP2}}\\ By Plancherel Theorem and by Lemma~\ref{lem:1}, one has $$ \Vert \Delta_j^{\alpha} f \Vert_{L^2}^2 = \int_{\mathbb{R}^2}|\varphi^{\alpha}_{j}(\xi)|^2 \vert \widehat{f}(\xi) \vert^2\mathrm{d}\xi \le C_0 \int |g_j(\xi)|^2 \vert \widehat{f}(\xi) \vert^2\mathrm{d}\xi\;, $$ and $$ \int_{\mathbb{R}^2} g_j^2(\xi) \vert \widehat{f}(\xi) \vert^2\mathrm{d}\xi \le \int_{\mathbb{R}^2} \left[(\varphi^ {\alpha}_{j-1})^2 + (\varphi^{\alpha}_j)^2 + (\varphi^{\alpha}_j)^2 \right] \vert \widehat{f}(\xi) \vert^2 \le \Vert \Delta_{j-1}^{\alpha} f \Vert_{L^2}^2 + \Vert \Delta_j^{\alpha} f \Vert_{L^2}^2 +\Vert \Delta_{j+1}^{\alpha} f \Vert_{L^2}^2\;, $$ where $g_j$ is defined by (\ref{e:compLPHyp}). Proposition~\ref{pro:caracLP2} is then a straightforward consequence of the following lemma : \begin{lemma} There exists some $C>0$ such that for any $j\ge 0$, one has \begin{equation}\label{e:equivL12} C^{-1} \sum_{(j_1,j_2)\in \Gamma_j(\alpha)} \Vert \phi_{j_1,j_2} \widehat{f} \Vert_{L^2}^2 \le \Vert g_j\widehat{f} \Vert_{L^2}^2 \le C\sum_{(j_1,j_2)\in \Gamma_j(\alpha)} \Vert \phi_{j_1,j_2} \widehat{f} \Vert_{L^2}^2 \;. \end{equation} \end{lemma} \begin{proof}Since $\mathrm{supp}(\phi_{j_1,j_2})\cap \mathrm{supp}(\phi_{m_1,m_2})=\emptyset$ if $\max(|m_1-j_1|,|m_2-j_2|)\ge 3$, the system $(\phi_{j_1,j_2} \widehat{f})$ is quasi--orthogonal. Hence, by Lemma~\ref{lem:qo}, there exists some $K>0$ such that $$ \left\Vert \sum_{(j_1,j_2)\in \Gamma_j(\alpha)} \phi_{j_1,j_2}\widehat{f} \right\Vert_{L^2}^2 \le K \sum_{(j_1,j_2)\in \Gamma_j(\alpha)} \Vert \phi_{j_1,j_2}\widehat{f} \Vert_{L^2}^2\;. $$ For the converse inequality, observe that each term $\langle \phi_{j_1,j_2}\widehat{f},\phi_{j'_1,j'_2}\widehat{f}\rangle=\int \phi_{j_1,j_2}(\xi)\phi_{j'_1,j'_2}(\xi) \vert f(\xi) \vert^2\mathrm{d} \xi$ is positive. Hence \begin{eqnarray*} \sum_{(j_1,j_2)\in \Gamma_j(\alpha)} \Vert \phi_{j_1,j_2}\widehat{f} \Vert_{L^2}^2 &\le& \sum_{(j_1,j_2)\in \Gamma_j(\alpha)} \langle \phi_{j_1,j_2}\widehat{f},\phi_{j_1,j_2}\widehat{f}\rangle_{L^2} + \sum_{(j_1,j_2)\neq (j'_1,j'_2)\in \Gamma_j(\alpha)} \langle \phi_{j_1,j_2}\widehat{f},\phi_{j'_1,j'_2}\widehat{f}\rangle_{L^2}\\ &=& \left\Vert \sum_{(j_1,j_2)\in \Gamma_j(\alpha)} \phi_{j_1,j_2}\widehat{f} \right\Vert_{L^2}^2\;. \end{eqnarray*} \end{proof} \subsubsection{Proof of the hyperbolic wavelet characterization of anisotropic Besov spaces} Let us first consider the general case where $(p,q)\in (0,+\infty]^2$,$\beta,s \in\mathbb{R}$ and $\alpha=(\alpha_1,\alpha_2)$ a fixed anisotropy. Intermediate spaces $\mathcal{E}^{s,\alpha}_{p,q,|\log|^\beta}(\mathbb{R}^2)$ are defined as the collection of functions $f$ of $\mathcal{S}'(\mathbb{R}^2)$ such as $$ \sum_{j \ge 0} j^{-\beta q} 2^{jsq} \sum_{(j_1,j_2) \in \Gamma_j(\alpha)} \Vert \Delta_{j_1,j_2}f \Vert_p^q <+\infty. $$ A norm on $\mathcal{E}^{s,\alpha}_{p,q,|\log|^\beta}(\mathbb{R}^2)$ is defined as follows \[ \|f\|_{\mathcal{E}^{s,\alpha}_{p,q,|\log|^\beta}}=\left( \sum_{j \ge 0} j^{-\beta q} 2^{jsq} \sum_{(j_1,j_2) \in \Gamma_j(\alpha)} \Vert \Delta_{j_1,j_2}f \Vert_p^q \right)^{1/q}\; \] such that the embeddings \begin{itemize} \item if $q<\infty$ \[ \mathcal{E}^{s,\alpha}_{p,q,|\log|^{\beta-\max(1/p-1,0)-\max(1-1/q,0)}}(\mathbb{R}^2)\hookrightarrow B^{s,\alpha}_{p,q,|\log|^\beta}(\mathbb{R}^2)\hookrightarrow\mathcal{E}^{s,\alpha}_{p,q,|\log|^{\beta+1/q}}(\mathbb{R}^2)\;. \] \item if $q=\infty$ \[ \mathcal{E}^{s,\alpha}_{p,\infty,|\log|^{\beta-\max(1/p-1,0)-1}}(\mathbb{R}^2)\hookrightarrow B^{s,\alpha}_{p,\infty,|\log|^\beta}(\mathbb{R}^2)\hookrightarrow\mathcal{E}^{s,\alpha}_{p,q,|\log|^{\beta}}(\mathbb{R}^2)\;. \] \end{itemize} are an exact rewriting of Theorem~\ref{th:LPChar}. In the special case where $p=2$, we proved in Proposition~\ref{pro:caracLP2} that $H^{s,\alpha}_{|\log|^\beta}(\mathbb{R}^2)=B^{s,\alpha}_{2,2,|\log|^\beta}(\mathbb{R}^2)$ and $\mathcal{E}^{s,\alpha}_{2,2,|\log|^{\beta}}(\mathbb{R}^2)$ coincide. In the following proposition, an hyperbolic wavelet characterization of spaces $\mathcal{E}^{s,\alpha}_{p,q}(\mathbb{R}^2)$ is given. Combining Proposition~\ref{pro:spacesEsp}, Theorems~\ref{th:LPChar} and~\ref{th:caracLP2} directly implies Theorems~\ref{th:WCBesov} and~\ref{th:WCBesov2}. \begin{proposition}\label{pro:spacesEsp} Let $(p,q)\in (0,+\infty]^2$, $s,\beta\in\mathbb{R}^2$. The following assertions are equivalent \begin{enumerate} \item $f \in \mathcal{E}^{s,\alpha}_{p,q,|\log|^\beta}(\mathbb{R}^2)$ \item $\left( \sum_{j \ge 0} j^{-\beta q} 2^{jsq} \sum_{(j_1,j_2) \in \Gamma_j} 2^{-(j_1+j_2)q/p} \left( \sum_{(k_1,k_2) \in \mathbb{Z}^2} \vert c_{j_1,j_2,k_1,k_2} \vert^p \right)^{\frac{q}{p}}\right)^{\frac{1}{q}} <+\infty$. \item $\left( \sum_{(j_1,j_2) \in \mathbb{N}_0^2} \left(\max(\frac{j_1}{\alpha_1}, \frac{j_2}{\alpha_2})\right)^{-\beta q} 2^{\left(\max(\frac{j_1}{\alpha_1}, \frac{j_2}{\alpha_2})s-\frac{j_1+j_2}{p}\right)q} \left( \sum_{(k_1,k_2) \in \mathbb{Z}^2} \vert c_{j_1,j_2,k_1,k_2} \vert^p \right)^{\frac{q}{p}} \right)^{\frac{1}{q}} <+\infty$. \end{enumerate} \end{proposition} Let us prove Proposition~\ref{pro:spacesEsp}. The equivalence between assertions $(2)$ and $(3)$ holds since for any $(j_1,j_2)\in \Gamma_j(\alpha)$, one has $\max (\frac{j_1}{\alpha_1}, \frac{j_2}{\alpha_2})+2-2\leq j \leq \max (\frac{j_1}{\alpha_1}, \frac{j_2}{\alpha_2})+2$ and $\cup \Gamma_j = \mathbb{N}_0^2$. The crucial point is the equivalence between assertions $(1)$ and $(2)$ .\\ \noindent{\bf Proof of implication $(1) \Rightarrow (2)$ of Proposition~\ref{pro:spacesEsp}}\\ The proof of this implication relies on the following sampling lemma which is a adaptation of Lemma~2.4 of~\cite{frazier:jawerth:1985} in the case of rectangular support. \begin{lemma}\label{lem:sample} Let $p\in (0,+\infty]$ and $j=(j_1,j_2)\in\mathbb{N}_0^2$. Suppose $g\in\mathcal{S}'(\mathbb{R}^2)$ and $\mathrm{supp}(\widehat{g})\subset\{\xi,\,|\xi_1|\leq 2^{j_1+1}\mbox{ and }|\xi_2|\leq 2^{j_2+1}\}$. Then there exists $C>0$ such that $$ \left( \sum_{(k_1,k_2) \in \mathbb{Z}^2} 2^{-(j_1+j_2)} \left\vert g\left(\frac{k_1}{2^{j_1}}, \frac{k_2}{2^{j_2}}\right) \right\vert^p \right)^{1/p} \le C \Vert g \Vert_{L^p}\;. $$ \end{lemma} \begin{proof} Let $\psi\in\mathcal{S}(\mathbb{R}^2)$ be such that $\mathrm{supp}(\widehat{\psi})\subset \{\xi,\,\max(|\xi_1|,|\xi_2|)\leq \pi\}$ and $\widehat{\psi}\equiv 1$ on $[-2,2]^2$. Set $\psi_j(x)=2^{j_1+j_2}\psi(2^{j_1}x_1,2^{j_2}x_2)$. One has $\widehat{\psi_j}\equiv 1$ on $[-2^{j_1+1},2^{j_1+1}]\times [-2^{j_2+1},2^{j_2+1}]$. By assumption $\mathrm{supp}(\widehat{g})\subset[-2^{j_1+1},2^{j_1+1}]\times [-2^{j_2+1},2^{j_2+1}]$, so that for any $x=(x_1,x_2)\in\mathbb{R}^2$ and any fixed $y=(y_1,y_2)\in\mathbb{R}^2$ \[ g(x+y)=(\psi_j\star g)(x+y)=(2\pi)^{-2}\int_{\xi_1=-2^{j_1+1}}^{2^{j_1+1}}\int_{\xi_2=-2^{j_2+1}}^{2^{j_2+1}}\widehat{\psi_j}(\xi)\widehat{g}(\xi)\mathrm{e}^{\mathrm{i} x\xi}\mathrm{e}^{\mathrm{i} y\xi}\mathrm{d} \xi\;. \] Denote $\widehat{h_j}$ the periodic extension of $\widehat{\psi_j}$ with period $2^{j_i+1}\pi$ for each variable $\xi_i$ ($i=1,2$). One has \begin{equation}\label{e:g} g(x+y)=(2\pi)^{-2}\int_{\xi_1=-2^{j_1+1}}^{2^{j_1+1}}\int_{\xi_2=-2^{j_2+1}}^{2^{j_2+1}}\left(\widehat{h_j}(\xi)\mathrm{e}^{\mathrm{i} x\xi}\right)\left(\widehat{g}(\xi)\mathrm{e}^{\mathrm{i} y\xi}\right)\mathrm{d} \xi\;. \end{equation} Using an expansion of $\widehat{h_j}\mathrm{e}^{\mathrm{i} x\xi}$ in two dimensional Fourier series, it comes \begin{eqnarray*} && \widehat{h_j}(\xi)\mathrm{e}^{\mathrm{i} x\xi} \\ & = & \sum_{(\ell_1,\ell_2)\in\mathbb{Z}^2}\left(\int_{\xi_1=-2^{j_1+1}\pi}^{2^{j_1+1}\pi}\int_{\xi_2=-2^{j_2+1}\pi}^{2^{j_2+1}\pi}\widehat{h_j}(\xi)\mathrm{e}^{\mathrm{i} x\xi}\mathrm{e}^{-\mathrm{i} 2^{-j_1}\ell_1\xi_1}\mathrm{e}^{-\mathrm{i} 2^{-j_2}\ell_2\xi_2}\right)\mathrm{e}^{\mathrm{i} 2^{-j_1}\ell_1\xi_1}\mathrm{e}^{\mathrm{i} 2^{-j_2}\ell_2\xi_2} \\ & =& \sum_{(\ell_1,\ell_2)\in\mathbb{Z}^2}\left(\int_{\xi_1=-2^{j_1+1}\pi}^{2^{j_1+1}\pi}\int_{\xi_2=-2^{j_2+1}\pi}^{2^{j_2+1}\pi}\widehat{\psi_j}(\xi)\mathrm{e}^{\mathrm{i} x\xi}\mathrm{e}^{-\mathrm{i} 2^{-j_1}\ell_1\xi_1}\mathrm{e}^{-\mathrm{i} 2^{-j_2}\ell_2\xi_2}\right)\mathrm{e}^{\mathrm{i} 2^{-j_1}\ell_1\xi_1}\mathrm{e}^{\mathrm{i} 2^{-j_2}\ell_2\xi_2} \\ &=& 2^{-(j_1+j_2)}\sum_{(\ell_1,\ell_2)\in\mathbb{Z}^2}\psi_j(x-2^{-j}\ell)\mathrm{e}^{\mathrm{i} 2^{-j_1}\ell_1\xi_1}\mathrm{e}^{\mathrm{i} 2^{-j_2}\ell_2\xi_2} \;, \end{eqnarray*} where for $j=(j_1,j_2)$ and $\ell=(\ell_1,\ell_2)$, the notation $2^{-j}\ell= (2^{-j_1}\ell_1,2^{-j_2}\ell_2)$ is used. Replacing $\widehat{h_j}(\xi)\mathrm{e}^{\mathrm{i} x\xi}$ with the last sum in Equation~(\ref{e:g}) yields that for any $x=(x_1,x_2)\in\mathbb{R}^2$ and any fixed $y=(y_1,y_2)\in\mathbb{R}^2$ \begin{eqnarray*} & & g(x+y)\\ &=&\frac{2^{-(j_1+j_2)}}{4\pi^2}\sum_{(\ell_1,\ell_2)\in\mathbb{Z}^2}\left(\int_{\xi_1=-2^{j_1+1}}^{2^{j_1+1}}\int_{\xi_2=-2^{j_2+1}}^{2^{j_2+1}}\psi_j(x-2^{-j}\ell)\mathrm{e}^{\mathrm{i} 2^{-j_1}\ell_1\xi_1}\mathrm{e}^{\mathrm{i} 2^{-j_2}\ell_2\xi_2}\left(\widehat{g}(\xi)\mathrm{e}^{\mathrm{i} y\xi}\right)\mathrm{d} \xi\right)\\ &=&2^{-(j_1+j_2)}\sum_{(\ell_1,\ell_2)\in\mathbb{Z}^2}g(2^{-j}\ell+y)\psi_j(x-2^{-j}\ell)\;. \end{eqnarray*} Hence for all $y\in \lambda_{j_1,j_2,k_1,k_2}=[2^{-j_1}k_1,2^{-j_1}(k_1+1))\times [2^{-j_2}k_2,2^{-j_2}(k_2+1))$ \begin{eqnarray*} \sup_{|z_1-2^{-j_1}k_1|\leq 2^{-j_1},|z_2-2^{-j_2}k_2|\leq 2^{-j_2}}|g(z)|&\leq& \sup_{|x_1|\leq 2^{-j_1}\sqrt{2},|x_2|\leq 2^{-j_2}\sqrt{2}}|g(x+y)|\\&\leq& 2^{-(j_1+j_2)}\sum_{(\ell_1,\ell_2)\in\mathbb{Z}^2}|g(2^{-j}\ell+y)|\cdot\sup_{\max(2^{j_1}|x_1|,2^{j_2}|x_2|)\leq \sqrt{2}}|\psi_j(x-2^{-j}\ell)|\\ &\leq &2^{-(j_1+j_2)}\sum_{(\ell_1,\ell_2)\in\mathbb{Z}^2}|g(2^{-j}\ell+y)|\cdot\frac{1}{(1+|\ell|)^M} \end{eqnarray*} where the last inequality follows from the fast decay of $\psi$. Take $M$ sufficiently large and use either the p triangular inequality either the H\"{o}lder inequality according whether $p\in (0,1)$ or $p\in [1,+\infty]$. Hence, one has \[ |g(2^{-j_1}k_1,2^{-j_2}k_2)|^p\leq\sup_{|z_1-2^{-j_1}k_1|\leq 2^{-j_1},|z_2-2^{-j_2}k_2|\leq 2^{-j_2}}|g(z)|^p\leq C 2^{-(j_1+j_2)}\sum_{(\ell_1,\ell_2)\in\mathbb{Z}^2}|g(2^{-j}\ell+y)|^p\cdot\frac{1}{(1+|\ell|)^{M'}}\;, \] for some $M'>1$. An integration over $y\in \lambda_{j_1,j_2,k_1,k_2}$ leads to \[ 2^{-(j_1+j_2)}|g(2^{-j_1}k_1,2^{-j_2}k_2)|^p\leq\sum_{(\ell_1,\ell_2)\in\mathbb{Z}^2}\frac{1}{(1+|\ell|)^{M'}}\int_{\lambda_{j_1,j_2,k_1,k_2}}|g(y)|^p \mathrm{d} y \] and a sum over $k\in \mathbb{Z}^2$ gives \[ \sum_k 2^{-(j_1+j_2)}|g(2^{-j_1}k_1,2^{-j_2}k_2)|^p\leq \sum_k\sum_{(\ell_1,\ell_2)\in\mathbb{Z}^2}\frac{1}{(1+|\ell|)^{3}}\int_{\lambda_{j_1,j_2,k_1,k_2}}|g(y)|^p \mathrm{d} y \] which ends the proof of Lemma~\ref{lem:sample}. \end{proof} Now, observe that $c_{j_1,j_2,k_1,k_2}=\Delta_{j_1,j_2}f(2^{-j_1}k_1, 2^{-j_2}k_2)$. By Lemma~\ref{lem:sample} applied to $g=\Delta_{j_1,j_2}f\in \mathcal{S}(\mathbb{R}^2)$, one has $$ \sum_{(k_1,k_2) \in \mathbb{Z}^2}|c_{j_1,j_2,k_1,k_2}|^p=\sum_{(k_1,k_2) \in \mathbb{Z}^2} \vert \Delta_{j_1,j_2}f(2^{-j_1}k_1, 2^{-j_2}k_2) \vert^ p \le C 2^{j_1} 2^{j_2}\Vert \Delta_{j_1,j_2} f \Vert_p^p\;, $$ which is the desired wavelet characterization. \\ \noindent{\bf Proof of implication $(2) \Rightarrow (1)$ of Proposition~\ref{pro:spacesEsp}}\\ To obtain the converse implication, the same approach as in the proof of Theorem~3.1 of~\cite{frazier:jawerth:1985} is followed. Since $\phi_{j_1,j_2}$ and $\psi_{m_1,m_2,k_{1},k_1}$ are both defined as a tensorial product, Lemma~3.3 of ~\cite{frazier:jawerth:1985} can be applied: there exists some $C>0$ such that for any $\alpha>0$ and for all $x=(x_1,x_2)\in\mathbb{R}^2$ one has \begin{equation}\label{e:ineqL33} |\phi_{j_1,j_2}\star \psi_{m_1,m_2,k_{1},k_1}(x)|\leq C \frac{2^{-(|j_1-m_1|+|j_2-m_2|)(M+3)}}{(1+2^{\inf(j_1,m_1)}|x_1-2^{-m_1}k_1|)^{\alpha}(1+2^{\inf(j_2,m_2)}|x_2-2^{-m_2}k_2|)^{\alpha}}\;, \end{equation} where $M$ denotes the number of vanishing moments of the wavelets. A lemma analogous to Lemma~3.4 of~\cite{frazier:jawerth:1985}~ is now proved: \begin{lemma}\label{lem:34} Let $p\in [1,+\infty]$, $\ell_1,\ell_2,m_1,m_2$ integers such that $\ell_1\leq m_1$ and $\ell_2\leq m_2$. We are also given some functions $g_{k_1,k_2}$ satisfying the following inequality for some $C>0$~: \begin{equation}\label{e:ineqL34a} \forall x=(x_1,x_2)\in\mathbb{R}^2,\,|g_{k_1,k_2}(x)|\leq \frac{C}{(1+2^{\ell_1}|x_1-2^{-m_1}k_1|)^2(1+2^{\ell_2}|x_2-2^{-m_2}k_2|)^2}\;. \end{equation} Set \[ F=\sum_{k=(k_1,k_2)\in\mathbb{Z}^2}d_{k_1,k_2}g_{k_1,k_2} \] Then \begin{equation}\label{e:ineqL34b} \|F\|_{L^p}\leq C2^{-(m_1+m_2)/p}2^{m_1-\ell_1}2^{m_2-\ell_2}\cdot\left(\sum_{k=(k_1,k_2)\in\mathbb{Z}^2}|d_{k_1,k_2}|^p\right)^{1/p}\;. \end{equation} \end{lemma} \begin{proof} By definition of the $L^p$--norm, one has~: \begin{eqnarray*} \|F\|_{L^p}^p&=&\int_{\mathbb{R}^2}\left|\sum_{k=(k_1,k_2)\in\mathbb{Z}^2}d_{k_1,k_2}g_{k_1,k_2}(x)\right|^p\mathrm{d} x\\ &\leq&\sum_{k'=(k'_1,k'_2)\in\mathbb{Z}^2}\int_{\lambda_{m_1,m_2,k'_1,k'_2}}\left|\sum_{k=(k_1,k_2)\in\mathbb{Z}^2}d_{k_1,k_2}g_{k_1,k_2}(x)\right|^p\mathrm{d} x\;, \end{eqnarray*} where the hyperbolic dyadic cube $\lambda_{m_1,m_2,k'_1,k'_2}$ are defined in~(\ref{e:dyadiccube}). Observe now that, by the usual triangular inequality and by inequality~(\ref{e:ineqL34a}), there exists some $C>0$ such that for any $(k_1,k_2)\in\mathbb{Z}^2$, $(k'_1,k'_2)\in\mathbb{Z}^2$ \[ \sup_{x\in \lambda_{m_1,m_2,k'_1,k'_2}}\left|\sum_{k=(k_1,k_2)\in\mathbb{Z}^2}d_{k_1,k_2}g_{k_1,k_2}(x)\right|\leq \sum_{k=(k_1,k_2)\in\mathbb{Z}^2}\frac{|d_{k_1,k_2}|}{\prod_{i=1,2}(1+2^{\ell_i}|2^{-m_i}k'_i-2^{-m_i}k_i|)^2} \] Hence one has \begin{eqnarray*} \|F\|_{L^p}^p&\leq&C 2^{-(m_1+m_2)}\sum_{(k'_1,k'_2)\in\mathbb{Z}^2} \left(\sum_{(k_1,k_2)\in\mathbb{Z}^2}\frac{|d_{k_1,k_2}|}{(1+2^{\ell_1-m_1}|k'_1-k_1|)^2(1+2^{\ell_2-m_2}|k'_2-k_2|)^2}\right)^p\;, \end{eqnarray*} Let us recall the usual convolution inequality, valid for any sequences $s,s'$ in $\ell^p(\mathbb{Z}^2)$ for $p\geq 1$, \[ \|s*s'\|_{\ell_p(\mathbb{Z}^2)}^p\leq \|s\|_{\ell_p(\mathbb{Z}^2)}^p\|s'\|_{\ell^1(\mathbb{Z}^2)}^p\;. \] Applied to $s=|d_{k_1,k_2}|$ and $s'=(1+2^{\ell_1-m_1}|k'_1-k_1|)^{-2}(1+2^{\ell_2-m_2}|k'_2-k_2|)^{-2}$, it gives \begin{eqnarray*} \|F\|_{L^p}^p&\leq &C 2^{-(m_1+m_2)}\left(\sum_{(k_1,k_2)\in\mathbb{Z}^2}|d_{k_1,k_2}|^p\right) \left(\sum_{(k'_1,k'_2)\in\mathbb{Z}^2}\frac{1}{(1+2^{\ell_1-m_1}|k'_1|)^2(1+2^{\ell_2-m_2}|k'_2|)^2}\right)^p\;, \end{eqnarray*} Recall now the classical result~: \[ \sum_{k'=(k'_1,k'_2)\in\mathbb{Z}^2}\frac{1}{(1+2^{\ell_1-m_1}|k'_1|)^2(1+2^{\ell_2-m_2}|k'_2|)^2}\leq C2^{m_1-\ell_1}2^{m_2-\ell_2} \] Hence \begin{eqnarray*} \|F\|_{L^p}^p&\leq& C 2^{-(m_1+m_2)} 2^{(m_1-\ell_1)p}2^{(m_2-\ell_2)p}\left(\sum_{k=(k_1,k_2)\in\mathbb{Z}^2}|d_{k_1,k_2}|^p\right) \\ & & \times \left(\sum_{k'=(k'_1,k'_2)\in\mathbb{Z}^2}\frac{1}{\prod_{i=1,2}(1+2^{\ell_i-m_i}|k'_i|)^2}\right)^p\;, \end{eqnarray*} which directly yields the required result. It ends the proof of Lemma~\ref{lem:34}. \end{proof} Let us now go back to Implication $(2) \Rightarrow (1)$ of Proposition~\ref{pro:spacesEsp}. Two cases are considered: $p\in (0,1)$ and $p\in [1,+\infty]$.\\ Let us first assume that $p\in (0,1)$. \\ We have to bound $\|\Delta_{j_1,j_2}f\|_{L^p}=\|\phi_{j_1,j_2}\star f\|_{L^p}$. Observe that \[ \phi_{j_1,j_2}\star f=\sum_{m_1,m_2}\sum_{k_1,k_2}c_{m_1,m_2,k_1,k_2}\left(\phi_{j_1,j_2}\star \psi_{m_1,m_2,k_1,k_2}\right) \] By the $p$--triangular inequality, it comes \[ \forall x=(x_1,x_2)\in\mathbb{R}^2,\,|\phi_{j_1,j_2}\star f(x)|^p\leq \sum_{m_1,m_2}\sum_{k_1,k_2}|c_{m_1,m_2,k_1,k_2}|^p\left|(\phi_{j_1,j_2}\star \psi_{m_1,m_2,k_1,k_2})(x)\right|^p \] By Inequality~(\ref{e:ineqL33}), for all $x=(x_1,x_2)\in\mathbb{R}^2$, one has \[ |\phi_{j_1,j_2}\star f(x)|^p\leq \sum_{m_1,m_2}\sum_{k_1,k_2}|c_{m_1,m_2,k_1,k_2}|^p\times\frac{2^{-p(|j_1-m_1|+|j_2-m_2|)(M+3)}}{(1+2^{\inf(j_1,m_1)}|x_1-2^{-m_1}k_1|)^{p\alpha}(1+2^{\inf(j_2,m_2)}|x_2-2^{-m_2}k_2|)^{p\alpha}} \] An integration over $\mathbb{R}^2$ implies that~: \[ \|\phi_{j_1,j_2}\star f(x)\|^p_{L^p}\leq \sum_{m_1,m_2}\sum_{k_1,k_2}|c_{m_1,m_2,k_1,k_2}|^p 2^{-p(|j_1-m_1|+|j_2-m_2|)(M+3)}\;. \] Hence \begin{eqnarray*} \|f\|_{\mathcal{E}^{s,\alpha}_{p,q,|\log|^\beta}}^q&=&\sum_{j_1,j_2}\left(\max(\frac{j_1}{\alpha_1},\frac{j_2}{\alpha_2})\right)^{-\beta q}2^{q s\max(\frac{j_1}{\alpha_1},\frac{j_2}{\alpha_2})}\|\phi_{j_1,j_2}\star f(x)\|^q_{L^p}\\ &\leq& \sum_{j_1,j_2}\left(\sum_{m_1,m_2}\|c_{m_1,m_2,\cdot,\cdot}\|^p_{\ell^p} 2^{-p(|j_1-m_1|+|j_2-m_2|)(M+3)}\left(\max(\frac{j_1}{\alpha_1},\frac{j_2}{\alpha_2})\right)^{-\beta p}2^{p s\max(\frac{j_1}{\alpha_1},\frac{j_2}{\alpha_2})}\right)^{q/p} \end{eqnarray*} Set for any $t\in\mathbb{R}$, $(t)_+=\max(t,0)$ and \[ \mathrm{sgn}(t)=\left\{\begin{array}{l}1\mbox{ if }t>0,\\0\mbox{ if }t=0,\\-1\mbox{ if }t<0.\end{array} \right. \] Observe now that for any integers $j,m$ \[ m-(m-j)_+\leq j\leq (j-m)_+ + m\;, \] and that for any integers $j_1,j_2,m_1,m_2$ \[ \frac{\max(\frac{m_1}{\alpha_1},\frac{m_2}{\alpha_1})}{1-\frac{\max(\frac{(m_1-j_1)_+}{\alpha_1},\frac{(m_2-j_2)_+}{\alpha_1})}{\max(\frac{m_1}{\alpha_1},\frac{m_2}{\alpha_1})}}\leq \max(\frac{j_1}{\alpha_1},\frac{j_2}{\alpha_1})\leq \max(\frac{m_1}{\alpha_1},\frac{m_2}{\alpha_1})\left[1+\max(\frac{(j_1-m_1)_+}{\alpha_1},\frac{(j_2-m_2)_+}{\alpha_1})\right]\;, \] (except in the case $m_1=m_2=0$ which can be treated separately). Hence \[ \|f\|_{\mathcal{E}^{s,\alpha}_{p,q,|\log|^\beta}}^q\leq \sum_{j_1,j_2}\left(\sum_{m_1,m_2} u_{m_1,m_2}v_{j_1-m_1,j_2-m_2}\cdot \right)^{q/p}\;, \] with \[ s_{m_1,m_2}=\left(\max(\frac{m_1}{\alpha_1},\frac{m_2}{\alpha_2})\right)^{-\beta p}2^{p s\max(\frac{m_1}{\alpha_1},\frac{m_2}{\alpha_2})}\|c_{m_1,m_2,\cdot,\cdot}\|^p_{\ell^p}\;, \] and \[ s'_{j_1,j_2}=2^{-p(|j_1|+|j_2|)(M+3)}[1+\max(\frac{(j_1)_+}{\alpha_1},\frac{(j_2)_+}{\alpha_2})]^{-\beta p}2^{\mathrm{sgn}(s) p s\max(\frac{(j_1)_+}{\alpha_1},\frac{(j_2)_+}{\alpha_2})} \] If $q/p>1$ Young's inequality can be applied, which states that for any sequences $s,s'$, \[ \|s*s'\|_{\ell^{q/p}(\mathbb{Z}^2)}\leq \|s\|_{\ell^{q/p}(\mathbb{Z}^2)}\|s'\|_{\ell^1(\mathbb{Z}^2)}\;, \] whereas if $q/p\leq 1$ the usual $(q/p)$--triangle inequality and the usual inequality $\|s*s'\|_{\ell^1(\mathbb{Z}^2)}\leq \|s\|_{\ell^1(\mathbb{Z}^2)}\|s'\|_{\ell^1(\mathbb{Z}^2)}$ valid for any sequence $s,s'$ are applied. In any case, the following inequality is obtained \begin{eqnarray*} &&\|f\|_{\mathcal{E}^{s,\alpha}_{p,q,|\log|^\beta}}^q\\ &\leq& \left(\sum_{m_1,m_2}\left(\max(\frac{m_1}{\alpha_1},\frac{m_2}{\alpha_2})\right)^{-\beta p}2^{q s\max(\frac{m_1}{\alpha_1},\frac{m_2}{\alpha_2})}\|c_{m_1,m_2,\cdot,\cdot}|^p_{\ell^p}\right)\\ &&\times\sum_{j_1,j_2}\left(2^{-p(j_1+j_2)(M+3)}\left(\max(\frac{j_1}{\alpha_1},\frac{j_2}{\alpha_2})\right)^{-\beta p}2^{p s\max(\frac{j_1}{\alpha_1},\frac{j_2}{\alpha_2})}\right)^{\max(q/p,1)}\;. \end{eqnarray*} If the wavelets have sufficiently vanishing moments, we get that \[ \|f\|_{\mathcal{E}^{s,\alpha}_{p,q,|\log|^\beta}}^q\leq C\left(\sum_{m_1,m_2}\left(\max(\frac{m_1}{\alpha_1},\frac{m_2}{\alpha_2})\right)^{-\beta p}2^{q s\max(\frac{m_1}{\alpha_1},\frac{m_2}{\alpha_2})}\|c_{m_1,m_2,\cdot,\cdot}|^p_{\ell^p}\right)\;, \] which is the required result.\\ \noindent We now consider the case $p\in [1,+\infty]$. In this case, observe that \[ \Delta_{j_1,j_2}f=\sum_{k_1,k_2}d_{k_1,k_2}g_{k_1,k_2} \] with \[ g_{k_1,k_2}=2^{(|j_1-m_1|+|j_2-m_2|)(M+3)}(\phi_{j_1,j_2}\star \psi_{m_1,m_2,k_1,k_2})\;, \] and \[ d_{k_1,k_2}=2^{-(|j_1-m_1|+|j_2-m_2|)(M+3)}c_{j_1,j_2,k_1,k_2}\;. \] We set $\ell_1=\inf(j_1,m_1)$ and $\ell_2=\inf(j_2,m_2)$. Lemma~\ref{lem:34} gives \[ \|\Delta_{j_1,j_2}f\|_{L^p}\leq C 2^{-p(|j_1-m_1|+|j_2-m_2|)(M+3)}2^{-(m_1+m_2)/p}\|c_{m_1,m_2,\cdot,\cdot}|^p_{\ell^p}2^{m_1-\ell_1}2^{m_2-\ell_2} \] Again two cases $q\leq 1$ and $q>1$ are distinguished and the same approach than in the case $p\in (0,1)$ is followed. It leads to the required conclusion. \subsection{Proof of Theorem~\ref{th:WL}}\label{s:proofWC} First a two--microlocal criterion is proved. \begin{proposition}\label{pro:WC2micro} \, \begin{enumerate} \item Assume that $f\in \mathcal{C}^{s,\alpha}(x_0)$. Then there exists some $C>0$ such that for any $(j_1,j_2,k_1,k_2)\in (\mathbb{N}\cup \{-1\})^2\times \mathbb{Z}^2$, \begin{equation}\label{e:WC2micro} |c_{j_1,j_2,k_1,k_2}|\leq C\min(2^{-\frac{j_1 s}{\alpha_1}}+\left|\frac{k_1}{2^{j_1}}-a\right|^{\frac{s}{\alpha_1}},2^{-\frac{j_2 s}{\alpha_2}}+\left|\frac{k_2}{2^{j_2}}-b\right|^{\frac{s}{\alpha_2}})\;. \end{equation} \item Conversely, assume that $f$ is uniformly H\"{o}lder and that~(\ref{e:WC2micro}) holds, then $f\in \mathcal{C}^{s,\alpha}_{|\log|^2}(x_0)$. \end{enumerate} \end{proposition} \begin{proof} Let us first assume that $f\in \mathcal{C}^{s,\alpha}(x_0)$ with $x_0=(a,b)$. Assume that $j_1\neq -1$ and $j_2\neq -1$. By definition of the hyperbolic wavelet coefficients one has \begin{eqnarray*} c_{j_1,j_2,k_1,k_2}&=&2^{j_1+j_2}\int_{\mathbb{R}^2} f(x_1,x_2)\psi(2^{j_1}x_1-k_1)\psi(2^{j_2}x_2-k_2)\mathrm{d} x_1\mathrm{d} x_2 \end{eqnarray*} Since $\psi$ admits at least one vanishing moment, the two following equalities hold \begin{equation}\label{e:eq1} c_{j_1,j_2,k_1,k_2}=2^{j_1+j_2}\int_{\mathbb{R}^2} (f(x_1,x_2)-P_{x_0}(a,x_2))\psi(2^{j_1}x_1-k_1)\psi(2^{j_2}x_2-k_2)\mathrm{d} x_1\mathrm{d} x_2 \end{equation} and \begin{equation}\label{e:eq2} c_{j_1,j_2,k_1,k_2}=2^{j_1+j_2}\int_{\mathbb{R}^2} (f(x_1,x_2)-P_{x_0}(x_1,b))\psi(2^{j_1}x_1-k_1)\psi(2^{j_2}x_2-k_2)\mathrm{d} x_1\mathrm{d} x_2 \end{equation} Equality~(\ref{e:eq1}) and the assumption $f\in\mathcal{C}^{s,\alpha}(x_0)$ imply that \begin{eqnarray*} |c_{j_1,j_2,k_1,k_2}|&\leq& 2^{j_1+j_2}\int |x_1-a|^s_\alpha|\psi(2^{j_1}x_1-k_1)\psi(2^{j_2}x_2-k_2)|\mathrm{d} x_1\mathrm{d} x_2\\ &\leq&2^{j_1+j_2}\int_{\mathbb{R}^2} \left(\left|x_1-\frac{k_1}{2^{j_1}}\right|^{s/\alpha_1}+\left|\frac{k_1}{2^{j_1}}-a|^{s/\alpha_1}\right)\right|\psi(2^{j_1}x_1-k_1)\psi(2^{j_2}x_2-k_2)|\mathrm{d} x_1\mathrm{d} x_2 \end{eqnarray*} We now set $u_1=2^{j_1}x_1-k_1$, $u_2=2^{j_2}x_2-k_2$ and deduce that \[ |c_{j_1,j_2,k_1,k_2}|\leq \left(2^{-\frac{j_1 s}{\alpha_1}}\int_{\mathbb{R}^2}|u_1|^{s/\alpha_1} |\psi(u_1)\psi(u_2)|\mathrm{d} u_1\mathrm{d} u_2+\left|\frac{k_1}{2^{j_1}}-a\right|^{s/\alpha_1}\int |\psi(u_1)\psi(u_2)|\mathrm{d} u_1\mathrm{d} u_2\right)\;. \] Hence for some $C$ depending only on $\psi$, $s$ and $\alpha$ one has \[ |c_{j_1,j_2,k_1,k_2}|\leq C(2^{-\frac{j_1 s}{\alpha_1}}+\left|\frac{k_1}{2^{j_1}}-a\right|^{s/\alpha_1}) \] A similar approach yields that \[ |c_{j_1,j_2,k_1,k_2}|\leq C(2^{-\frac{j_2 s}{\alpha_2}}+\left|\frac{k_2}{2^{j_2}}-b\right|^{s/\alpha_2}) \] This shows that (\ref{e:WC2micro}) can be read as a necessary condition for pointwise regularity of function $f$. Let us now prove the converse result. Assuming that~(\ref{e:WC2micro}) holds, the aim first consists in defining a polynomial approximation of $f$ at $x_0$. To that end, a Taylor expansion is used to investigate the differentiability of $f$ at $x_0$. Let us define $f_j$ as: \[ f_j=\sum_{(j_1,j_2)\in \Gamma_j(\alpha)}\sum_{(k_1,k_2)\in\mathbb{Z}^2}c_{j_1,j_2,k_1,k_2}\psi_{j_1,j_2,k_1,k_2}\;. \] where the notations are the same as in the proof of Proposition~\ref{pro:WCGlobal}. One has \begin{eqnarray*} |f_j(x)|&\leq& \sum_{(j_1,j_2)\in \Gamma_j}\sum_{(k_1,k_2)\in\mathbb{Z}^2}\frac{\min(2^{-j_1s/\alpha_1}+|\frac{k_1}{2^{j_1}}-a|^{s/\alpha_1},2^{-j_2 s/\alpha_2}+|\frac{k_2}{2^{j_2}}-a|^{s/\alpha_2})} {(1+|2^{j_1}x_1-k_1|)^N(1+|2^{j_2}x_2-k_2|)^N}\\ &\leq& \sum_{j_1\leq j}\sum_{k_1,k_2}\frac{2^{-j s}+|\frac{k_2}{2^{j}}-x_2|^{s/\alpha_2}+|x_2-b|^{s/\alpha_2}}{(1+|2^{j_1}x_1-k_1|)^N(1+|2^{j_2}x_2-k_2|)^N}+\sum_{j_2\leq j}\frac{2^{-js}+|\frac{k_1}{2^{j_1}}-x_1|^{s/\alpha_1}+|x_1-a|^{s/\alpha_2}}{(1+|2^{j_1}x_1-k_1|)^N(1+|2^{j_2}x_2-k_2|)^N} \end{eqnarray*} Then \begin{equation}\label{e:majofj1} |f_j(x)|\leq C(j2^{-js}+j|x_1-a|^{s/\alpha_1}+j|x_2-b|^{s/\alpha_2})\;. \end{equation} In the same way, if $\beta=(\beta_1,\beta_2)$, one has \[ |\partial^\beta f_j|\leq \sum_{(j_1,j_2)\in \Gamma_j}2^{j_1\beta_1+j_2\beta_2}\sum_{(k_1,k_2)\in\mathbb{Z}^2}\frac{\min(2^{-j_1s/\alpha_1}+|\frac{k_1}{2^{j_1}}-a|^{s/\alpha_1},2^{-j_2 s/\alpha_2}+|\frac{k_2}{2^{j_2}}-a|^{s/\alpha_2})} {(1+|2^{j_1}x_1-k_1|)^N(1+|2^{j_2}x_2-k_2|)^N}\;. \] Then \begin{equation}\label{e:majofj2} |\partial^\beta f_j(x)|\leq C2^{j(\beta_1\alpha_1+\beta_2\alpha_2)}(2^{-js}+|x_1-a|^{s/\alpha_1}+|x_2-b|^{s/\alpha_2})\;. \end{equation} So, the function $f$ is $\beta$--differentiable at $x_0$ provided that $\beta_1\alpha_1+\beta_2\alpha_2\leq s$. The Taylor polynomial of $f$ at $x_0$ is defined by \[ P_{j,x_0}(x)=\sum_{\beta_1\alpha_1+\beta_2\alpha_2\leq s}\frac{(x-x_0)^\beta}{\beta!}\partial^\beta f_j(x_0) \] and \[ P_{x_0}(x)=\sum_j P_{j,x_0}(x)\;. \] We shall now bound $|f(x)-P_{x_0}(x)|$ in the neighborhood of $x_0$. Recall that $f$ is assumed to be uniformly H\"{o}lder, namely there exists some $\varepsilon_0^*>0$ such that $f\in\mathcal{C}^{\varepsilon_0^*}(\mathbb{R}^2)$. The inclusions between H\"{o}lder spaces with different anisotropies (see \cite{triebel:2006}) leads to the existence of $\varepsilon_0$ such that $f\in\mathcal{C}^{\varepsilon_0,\alpha}(\mathbb{R}^2)$. Set $J_1=[\alpha J/\varepsilon_0]$. Observe that \[ |f(x)-P_{x_0}(x)|\leq \sum_{j\leq J}|f_j(x)-P_{j,x_0}(x)|+\sum_{j= J+1}^{J_1}|f_j(x)|+\sum_{j>J_1}|f_j(x)|+\sum_{j>J}|P_{j,x_0}(x)|\;. \] Let us now bound each term of the right hand side of this inequality. We first deal with the term corresponding to $j\leq J$. In this case we shall use an anisotropic version of Taylor inequality which can be found in~\cite{calderon:torchinsky:1977}, \cite{folland:stein:1982} and recalled in~\cite{benbraiek:benslimane:2011b}. It gives the existence of some $C>0$ such that \[ |f_j(x)-P_{j,x_0}(x)|\leq C\sum_{\beta_1+\beta_2\leq k+1,\,\alpha_1\beta_1+\alpha_2\beta_2>s}|x-x_0|^{\alpha_1\beta_1+\alpha_2\beta_2}_\alpha\sup_{z=(z_1,z_2)\in\mathbb{R}^{2}} |\partial^\beta f_j|\;. \] with $k=[\max(s/\alpha_1,s/\alpha_2)]$. The bound~(\ref{e:majofj2}) implies that there exists some $C>0$ such that \[ |f_j(x)-P_{j,x_0}(x)|\leq C \sum_{\beta_1+\beta_2\leq k+1,\,\alpha_1\beta_1+\alpha_2\beta_2>s}|x-x_0|^{\alpha_1\beta_1+\alpha_2\beta_2}_\alpha 2^{j(\beta_1\alpha_1+\beta_2\alpha_2)}(2^{-js}+|x_1-a|^{s/\alpha_1}+|x_2-b|^{s/\alpha_2}) \] Hence, \[ \sum_{j\leq J}|f_j(x)-P_{j,x_0}(x)|\leq C \sum_{\beta_1+\beta_2\leq k+1,\,\alpha_1\beta_1+\alpha_2\beta_2>s}|x-x_0|^{\alpha_1\beta_1+\alpha_2\beta_2}_\alpha(2^{J(\beta_1\alpha_1+\beta_2\alpha_2-s)}+ 2^{J(\beta_1\alpha_1+\beta_2\alpha_2)}|x-x_0|^s_\alpha)\;. \] Since $|x-x_0|_\alpha\leq 2^{-J}$ it comes \begin{equation}\label{e:ineq1} \sum_{j\leq J}|f_j(x)-P_{j,x_0}(x)|\leq C |x-x_0|^s_\alpha \end{equation} Let us now bound the sum $\sum_{j= J+1}^{J_1}|f_j(x)|$. By~(\ref{e:majofj1}) and the definition of $J_1$ which depends on $J$, one has \begin{equation}\label{e:ineq2} \sum_{j= J+1}^{J_1}|f_j(x)|\leq \sum_{j= J}^{J_1}(j2^{-js}+j|x-x_0|^s_\alpha)\leq J2^{-Js}+J^2|x-x_0|^s_\alpha\;. \end{equation} To bound the sum $\sum_{j>J_1}|f_j(x)|$ the uniform regularity of $f$ is used, leading to \begin{equation}\label{e:ineq3} \sum_{j>J_1}|f_j(x)|\leq C2^{-J_1\varepsilon_0}\leq C2^{-Js} \end{equation} the last equality following from the definition of $J_1$. Finally, by~(\ref{e:majofj2}), the sum $\sum_{j>J}|P_{j,x_0}(x)|$ can be bounded. Indeed, for some $C>0$, one has \[ \sum_{j>J}|P_{j,x_0}(x)|\leq \sum_{\beta_1\alpha_1+\beta_2\alpha_2<s}\frac{|(x-x_0)^\beta|}{\beta!}\sum_{j>J}|\partial^\beta f_j(x_0)|\leq C \sum_{\beta_1\alpha_1+\beta_2\alpha_2<s}\frac{|x_1-a|^{\beta_1}|x_2-b|^{\beta_2}}{\beta!}\sum_{j>J}2^{j(\beta_1\alpha_1+\beta_2\alpha_2-s)} \] Since $|x_1-a|\leq |x-x_0|_\alpha^{\alpha_1}\leq 2^{-J\alpha_1}$ and $|x_2-b|\leq |x-x_0|_\alpha^{\alpha_2}\leq 2^{-J\alpha_2}$ it comes \begin{equation}\label{e:ineq4} \sum_{j>J}|P_{j,x_0}(x)|\leq C \sum_{\beta_1\alpha_1+\beta_2\alpha_2<s}2^{-J(\beta_1\alpha_1+\beta_2\alpha_2)}\sum_{j>J}2^{j(\beta_1\alpha_1+\beta_2\alpha_2-s)}\leq C2^{-Js}\;. \end{equation} Finally, Inequalities~(\ref{e:ineq1}), (\ref{e:ineq2}), (\ref{e:ineq3}) and (\ref{e:ineq4}) yield that $f\in\mathcal{C}^{s,\alpha}_{|\log|^2}(x_0)$. \end{proof} Theorem~\ref{th:WL} is a straightforward consequence of the two--microlocal criterion and of the following lemma: \begin{lemma} The two following properties are equivalent: \begin{enumerate}[(i)] \item Inequality~(\ref{e:WC2micro}) holds. \item Inequality~(\ref{e:WL}) holds. \end{enumerate} \end{lemma} \begin{proof} Assume that~(\ref{e:WC2micro}) holds. If $\lambda'\subset 3\lambda_{j_1,j_2}(x_0)$, then \[ j'_1\geq j_1,j'_2\geq j_2\;, \] and \[ |\frac{k'_1}{2^{j'_1}}-a|\leq 2.2^{-j'_1}\mbox{ and }|\frac{k'_2}{2^{j'_2}}-b|\leq 2.2^{-j'_2}\;. \] Condition~(\ref{e:WC2micro}) implies \[ |c_{\lambda'}|\leq \min (2^{-\frac{j_1s}{\alpha_1}},2^{-\frac{j_2s}{\alpha_2}})=2^{-\max(\frac{j_1}{\alpha_1},\frac{j_2}{\alpha_2})s}\;. \] Conversely, assume that~(\ref{e:WL}) holds. Let $\lambda'=\lambda(j'_1,j'_2,k'_1,k'_2)$ an hyperbolic dyadic cube. Set \[ j_1=\sup\{\ell_1,\,2^{-j'_1}+|\frac{k'_1}{2^{j'_1}}-a|\leq 2^{-\ell_1}\} \] and \[ j_2=\sup\{\ell_2,\,2^{-j'_2}+|\frac{k'_2}{2^{j'_2}}-b|\leq 2^{-\ell_2}\} \] We have $\lambda'\subset3\lambda_{j_1,j_2}(x_0)$. Since~(\ref{e:WL}) holds one has \[ |c_{\lambda'}|\leq \min(2^{-\frac{j_1 s}{\alpha_1}},2^{-\frac{j_2 s}{\alpha_2}})\leq C\min(2^{-\frac{j'_1 s}{\alpha_1}}+\left|\frac{k'_1}{2^{j'_1}}-a\right|^{s/\alpha_1},2^{-\frac{j'_2 s}{\alpha_2}}+\left|\frac{k'_2}{2^{j'_2}}-b\right|^{\frac{s}{\alpha_2}})\;, \] that is ~\ref{e:WC2micro} holds. \end{proof} \subsection{Proof of Theorem~\ref{th:multform}}\label{s:proofmultimulti} The proof of Theorem~\ref{th:multform} is based on the two following lemmas, analogous to Propositions~7 and~8 of~\cite{jaffard:2004}: \begin{lemma}\label{lem1} Set $\alpha=(a,2-a)$ and define \[ G(H,\alpha)=\{x\in \mathbb{R}^2,\,f\not\in\mathcal{C}^{H,\alpha}_{|\log|^2}(x)\}\;. \] Let $p>0$ and $s\in (0,\omega(p,\alpha)/p]$. Then for any $H\geq s-2/p$ \[ \mathrm{dim}_H(G(H,\alpha))\leq Hp-sp+2\;. \] If $H< s-2/p$, $\mathrm{dim}_H(G(H,\alpha))=\emptyset$. \end{lemma} \begin{lemma}\label{lem2} Set $\alpha=(a,2-a)$ and define \[ B(H,\alpha)=\{x\in \mathbb{R}^2,\,f\in\mathcal{C}^{H,\alpha}(x)\}\;. \] Let $p<0$ and $s\in (0,\omega(p,\alpha)/p]$. Then \[ \mathrm{dim}_H(B(H,\alpha))\leq \mathrm{dim}_P(B(H,\alpha)) \leq Hp-sp+2\;. \] \end{lemma} The proof of Lemma~\ref{lem1} in the case $H\geq s-2/p$ is exactly the same as this of Proposition~7 of \cite{jaffard:2004}, except that the set $G_{j,H}$ are replaced with the sets \[ G(j,H,\alpha)=\{\lambda=\lambda(j_1,j_2,k_1,k_2),\, (j_1,j_2)\in \Gamma_j(\alpha),|d_\lambda|\geq 2^{-jHp}\}\;. \] Lemma~\ref{lem1} in the case $H< s-2/p$, comes from the hyperbolic wavelet characterization of anisotropic Besov spaces stated in Theorem~\ref{th:WCBesov} and the Sobolev embeddings which can be proved in the anisotropic case as in the isotropic one (see~\cite{triebel:2006}). The proof of Lemma~\ref{lem2} is exactly the same as this of Proposition~8 of \cite{jaffard:2004}, except that the set $B_{H}$ are replaced with the sets $B(H,\alpha)$. Lemmas~\ref{lem1} and \ref{lem2} then imply Theorem~\ref{th:multform}, since for any $\alpha=(\alpha_1,\alpha_2)$ such that $\alpha_1+\alpha_2=2$ one has \[ E(H,\alpha)\subset \left(\cap_{H'>H}G(H',\alpha)\right)\cap\left(\cup_{H'<H}B(H',\alpha)\right)\;. \] {\bf Acknowledgements.} We warmly thank Florent Autin and Jean Marc Freyermuth for many stimulating discussions about applications of non parametric statistics to the analysis of anisotropic textures as well as Laurent Duval for giving us very interesting additional references about hyperbolic wavelet analysis. \bibliographystyle{siam}
\section{Introduction} Ultra-cool dwarfs (UCDs) are objects with spectral type later than M7, comprising the lowest mass stars and brown dwarfs. Twelve of these (out of $\sim$200 observed) have been found to be intense sources of radio emissions with spectral luminosities typically of order a $\mathrm{MW\;Hz^{-1}}$ \citep{berger:2001aa, berger:2009aa, berger:2006aa, hallinan:2006aa,hallinan:2007aa, hallinan:2008aa,antonova:2007aa,antonova:2008aa, phan-bao:2007aa,mclean:2011aa,mclean:2012aa, route:2012aa}. The unpolarised component probably includes synchrotron emission, but the radio emission from a number of these, which have fast ($\mathbf{\sim}$2~h) rotation periods and strong ($\sim$$0.1$~T) magnetic fields, has been shown to be highly circularly polarised and modulated at the bodies' rotation periods \citep[e.g.][]{hallinan:2006aa}. The polarised nature of the emission implicates the electron cyclotron maser instability (CMI) as the source mechanism \citep{wu79a, treumann:2006aa}. \cite{yu:2011aa} modelled the radio emission, assuming it is generated by unaccelerated precipitating hot plasma trapped on coronal loops. However, this process would be likely limited to short-lived flare-type events, does not directly result in increased radio power with faster rotation as is observed \citep{mclean:2012aa}, and the CMI is known to be generated at much greater efficiency when electrons are accelerated down magnetic field lines by field-aligned voltages \citep{treumann:2006aa}. The presence of sustained CMI-generated radio emissions is thus instead strongly suggestive of the existence of quasi-steady auroral magnetic field-aligned currents. Indeed, all CMI source regions observed in situ at planets in the solar system exhibit accelerated electron populations, and radio emissions generated by the CMI have been extensively shown to be closely associated with auroral emission, caused when downward-precipitating electrons impact the atmosphere \citep{treumann:2006aa, clarke09a,lamy09a,lamy:2010aa, nichols10a,nichols10b}. \cite{yu:2012aa} have recently developed their analysis to consider the results of an accelerated population of electrons, and their simulations indicate that such populations could produce the observed UCD emissions. \\ Field-aligned currents arise from a divergence in field-perpendicular currents, which are, through $\mathbf{E}=-\mathbf{v}\times\mathbf{B}$, driven by plasma velocities relative to the neutrals in the conducting outer layer of the atmosphere. Thus, strong field-aligned currents flow when there is a sharp gradient in this departure from corotation, giving rise to a strong divergence in the field-perpendicular currents. We may therefore infer from the observation of CMI-generated radio emissions from UCDs that such angular velocity gradients with auroral currents exists in the magnetospheres of these objects. At planets in the solar system angular velocity gradients occur near the boundary between open and closed field lines, as at Earth and Saturn, or are due to centrifugally-driven outflow of internally-generated plasma, as at Jupiter. \cite{schrijver:2009aa} presented the hypothesis that the radio emission from UCDs could result from a Jupiter-like current system on the basis of a simple dimensional scaling relation, but did not consider quantitatively the magnetosphere-ionosphere (M-I) coupling current system. Here we have developed a simple axisymmetric model of the currents arising from departure from rigid corotation in the magnetospheres of UCDs, based on that used previously to study Jupiter's auroral oval \citep{hill79,hill01, cowley01,cowley02,nichols03,nichols04, nichols05,cowley05a,nichols11b} and to estimate the radio luminosity of fast-rotating Jupiter-like exoplanets \citep{nichols11a,nichols:2012aa}. We explain the properties of the radio emission from UCDs by showing that it would arise from the electric currents resulting from an angular velocity shear in the fast-rotating magnetic field and plasma, i.e.\ by an extremely powerful analogue of the process which causes Jupiter's or Saturn's auroras. Such a velocity gradient indicates that these bodies interact significantly with their space environment, resulting in intense auroral emissions.\\ \section{Theoretical background} \label{sec:theory} \subsection{Basic assumptions} \label{sec:assumps} For the purposes of development of the model, we assume that a jovian picture of an ultra-cool dwarf (UCD) is appropriate, i.e.\ that the UCD's magnetic field is generated deep within its interior, such that the external field generated by the body is mainly dipolar, and that a non-conducting atmosphere surrounds the body, the upper portion of which is maintained to some degree conductive by photon and/or corpuscular impact from the outside. The atmospheres of cool UCDs are not expected to be strongly ionised, although it has been suggested that dust clouds may induce some degree of ionisation \citep{helling:2011aa,helling:2011ab}. At solar system planets a number of sources of conductivity are known besides solar X-ray and EUV radiation, such as galactic cosmic rays and auroral electron precipitation \citep{rycroft:2008aa}. The latter is particularly important, since it results in a positive feedback which greatly increases the intensity of the auroral currents. At Jupiter, for example, modeling work indicates that the feedback from auroral precipitation increases the conductance from a residual value of $\sim$0.01~mho (where 1 mho $\equiv$ 1 siemens) generated e.g.\ by solar X-ray and EUV irradiation to values of up to 10~mho \citep{strobel83,millward02}, such that the feedback plays a dominant role in influencing the location and intensity of the M-I coupling currents \citep{nichols04}. Thus we suggest that given some small initial conductivity generated by a source mentioned above, the auroral current system would amplify itself via the feedback from the auroral precipitation. The exact details of such feedback requires ionospheric modeling that is beyond the scope of the present paper, but suffice it to say that due to this effect the low temperatures of UCDs do not necessarily prohibit the flow of strong M-I coupling currents. In the absence of such an ionospheric model for UCDs, in the present work we simply take the Pedersen conductance $\Sigma_P$ to be a constant, discussed further below.\\ A principal assumption is that the spin and magnetic axes are roughly co-aligned, since this is the configuration that has been most studied for bodies in our solar system, although we note that it is probable that only $\sim$50\% of UCDs exhibit this polarity. With reversed field polarity, the associated M-I coupling current system described below would also be reversed, such that the upward current, which in Fig. 1 is strongly peaked at 15$^\circ{}$, would instead be widely distributed at high latitudes, and the strong downward currents would be easily carried by upward-going ionospheric electrons, leading to little auroral and radio emissions. We also assume approximate axisymmetry about the magnetic axis, such that the flows in the ionosphere are considered to be wholly azimuthal about that axis, and can thus be described in terms of departure from corotation with the neutral atmosphere as a function of colatitude $\theta_i$ from the magnetic axis. Such a simplification seems appropriate for representing to lowest order the dynamics of a fast rotating magnetosphere, although we note that some processes related to an open magnetosphere will consequently be excluded from the model, such as `cusp-like' or `substorm-like' processes which would be ordered in azimuth with respect to the object's velocity through the surrounding medium. We further assume for simplicity that departures of the upper neutral atmosphere from rigid corotation due to ion-neutral drag also take the form of winds that are approximately axisymmetric about the magnetic axis, thus neglecting the effect of the Coriolis force on such motions in cases where the spin and magnetic axes are significantly inclined.\\ It has been suggested by \cite{kuznetsov:2012aa} that the short duration of the spikes in the CMI-induced radio emissions is indicative of an active sector in the magnetosphere of a UCD. We suggest, however, that the observed short duty cycles of $\leq$0.15 for UCD radio emissions \citep{hallinan:2006aa} are not inconsistent with an essentially axisymmetric field-aligned current distribution. Jupiter's main auroral oval is to first order described by an axisymmetric annulus of isotropic emission $\sim$$1^\circ{}$ wide located $\sim$$15^\circ{}$ from the magnetic pole, which is tilted by $\sim$10$^\circ{}$ from the spin axis \citep{grodent03b, clarke04, nichols09b}. Therefore, as the dipole axis cones around the spin axis, from a viewpoint near the planet's spin equatorial plane over half of the auroral oval is visible over roughly a third of the planet's rotation period \citep{nichols07}. The associated radio emission, however, is strongly beamed, into $\sim1.6$~sr in the case of the HOM and DAM emissions \citep{zarka04a}. Hence, the radio emission peaks strongly as the beam rotates into view, and the FWHM of the occurrence probability distribution of the `Source A' DAM emission (i.e.\ that used to determine Jupiter's System-III rotation period of $\sim$9~h 55~min) is $\sim$$50^\circ{}$, i.e.\ an apparent duty cycle of $\sim$$0.14$ \citep{higgins:1996aa}. This is consistent with the duty cycles of $\leq0.15$ for UCDs. Having said this, it is important to recognise that axisymmetry is simply a lowest order approximation. Saturn's radio emissions, for example, are observed from all local times but are brightest in the morning sector \citep{lamy09a}, despite the internal planetary magnetic field being axisymmetric to within detection limits \citep{burton:2010aa}. Finally, as has also been suggested by \cite{berger:2009aa}, it is probable that as is observed at Jupiter the magnetic poles may not lie on respective antipodes, due to the non-dipolar terms in the internal field, such that the radio emission from the two hemispheres would not necessarily be expected to be observed 180$^\circ$ apart in phase. \\ \subsection{Current system equations} \label{sec:currents} We consider three angular velocities in the computation of the M-I coupling currents. The first is the angular velocity of the deep interior of the UCD $\Omega_{UCD}$, taken initially to be $8.7\times10^{-4}\mathrm{\;rad\;s^{-1}}$ in conformity with the observed periods of $\sim$2~h for radio-bright UCDs. The second is the angular velocity of the plasma above the ionosphere $\omega$, taken to be constant on each flux tube, and the third is the angular velocity of the atmospheric neutrals in the Pedersen layer of the ionosphere $\Omega_{UCD}^*$, which is expected to lie somewhere between $\Omega_{UCD}$ and $\omega$ due to the frictional drag of ion-neutral collisions \citep{huang89, millward05}. In such a case we have \begin{equation} (\Omega_{UCD}-\Omega_{UCD}^*)=k(\Omega_{UCD}-\omega)\;\;, \label{eq:slip} \end{equation} \noindent where $0<k<1$. The actual value of $k$ is somewhat uncertain for Jupiter, with modelling work indicating a value of $\sim$0.5 \citep{millward05}, which is employed in the previous jovian modelling work discussed and also now adopted here, although the results are not expected to be sensitively dependent on reasonable choices. We should note that in the following work we consider only sub-rotation of the plasma with respect to the UCD. For Jupiter, super-rotation of the magnetospheric plasma may occur following strong solar wind-induced compressions of the magnetosphere, with a corresponding reversal of the current system described below \citep{hanlon04b,cowley07, nichols07}, but in the steady state the plasma sub-rotates, associated with the transfer of angular momentum from the atmosphere to the external medium. \\ The equatorward-directed height integrated ionospheric Pedersen current is then given by \begin{equation} i_P = \Sigma_P E_i=\Sigma_P\rho_i(\Omega_{UCD}^* - \omega)B_i\;\;, \label{eq:ip1} \end{equation} \noindent where $E_i$ is the electric field in the rest frame of the neutral atmosphere, $\rho_i = R_{UCD} \sin\theta_i$ is the distance from the spin axis (where we assume initially the UCD has a radius $R_{UCD}$ equal to Jupiter's equatorial 1-bar level radius $R_J = 71,373$~km since all bodies of roughly solar composition from giant planets through to the very lowest mass stars have radii roughly similar to Jupiter). Employing equation~\ref{eq:slip} in equation~\ref{eq:ip1} and introducing the `effective' Pedersen conductance $\Sigma_P^*$, reduced from the true value by a factor of $(1-k)$ owing to the slippage of the neutral atmosphere from rigid corotation, yields \begin{equation} i_P = \Sigma_P^*\rho_i(\Omega_{UCD}-\omega)B_i\;\;. \label{eq:ip2} \end{equation} \noindent As mentioned above, we simply take a constant value for $\Sigma_P^*$, initially 0.5~mho, which is similar to the values employed in previous Jupiter modelling works, and, as shown below, yields radio power values in agreement with observations of UCDs. However, in the results below we also examine the effect of different values of this parameter. The total Pedersen current at a given co-latitude is then found by integrating in azimuth, such that \begin{equation} I_P = 2\pi\rho_ii_P = 2\pi \Sigma_P^* \rho_i^2(\Omega_{UCD}-\omega)B_i\;\;, \label{eq:totip} \end{equation} \noindent the divergence of which gives the field-aligned current density just above the ionosphere, i.e. for the northern hemisphere\ \begin{equation} j_{\|i} = -\frac{1}{2\pi \rho_i^2 \sin \theta_i}\frac{dI_p}{d\theta_i}\;\;. \label{eq:jpari} \end{equation} \subsection{Field-aligned acceleration and energy flux} \label{sec:fav} The upward-directed field-aligned current computed in the previous section can not, in general, be carried simply by precipitating magnetospheric electrons alone, and must be driven by downward acceleration of those electrons by a field-aligned voltage. The maximum field-aligned current density that can be carried by an unaccelerated isotropic Maxwellian population is \begin{equation} j_{\|i0} = en\left(\frac{W_{th}}{2\pi m_e}\right)^{1/2}\;\;, \label{eq:jpari0} \end{equation} \noindent where $e$, $m_e$, $n$ and $W_{th}$ are the charge, mass, number density and thermal energy of the electron source population, respectively, the latter being equal to equal to $k_BT$, where $k_B$ is Boltzmann's constant and $T$ is the temperature. In common with previous works we use the values for the high latitude hot plasma at Jupiter, i.e.\ $n=0.01~\mathrm{cm}^{-3}$ and $W_{th}=2.5$~keV \citep{scudder81}, which, as shown below yield results consistent with observations, but we also examine the effects of wide ranges of values for these parameters. The corresponding unaccelerated kinetic energy flux is \begin{equation} E_{f0} = 2enW_{th}\left(\frac{W_{th}}{2\pi m_e}\right)^{1/2}\;\;. \label{eq:ef0} \end{equation} \noindent For field-aligned current densities larger than $j_{\|i0}$, a field-aligned voltage is required. For Earth, the current-voltage relation derived using the kinetic theory of \cite{knight73} is applicable, but for more powerful systems in which either the source electron population is very energetic initially, or becomes so following acceleration by voltages comparable to or exceeding the electron rest mass energy ($\sim$511~keV), Cowley's (2006) \nocite{cowley06b} relativistic extension to Knight's theory becomes appropriate. In the results shown in this paper it is clear that the relativistic theory is required in the case of M-I coupling current systems at UCDs, such that we employ the relativistic current-voltage relation given by \begin{equation} \left(\frac{\ensuremath{j_{\|i}}}{\ensuremath{j_{\|i\circ}}}\right)= 1 + \left(\frac{e\Phi}{W_{th}}\right)+ \frac{\left(\frac{e\Phi}{W_{th}}\right)^2}{2\left[\left(\frac{m_ec^2}{W_{th}}\right)+1\right]}\;\;, \label{eq:phi} \end{equation} \noindent where $c$ is the speed of light and $\Phi$ is the minimum voltage required to drive the current $j_{\|i}$ at ionospheric altitude, applicable in the limit that the accelerator is located at high altitude along the field line (where $B \ll B_i$). The corresponding precipitating electron energy flux is given by \begin{equation} \left(\frac{E_f}{E_{f0}}\right) = 1 + \left(\frac{e\Phi}{W_{th}}\right) + \frac{1}{2}\left(\frac{e\Phi}{W_{th}}\right)^2 + \frac{\left(\frac{e\Phi}{W_{th}}\right)^3} {2\left[2\left(\frac{m_ec^2}{W_{th}}\right)+3\right]}\;\;. \label{eq:ef} \end{equation} \noindent Equations~\ref{eq:jpari0}-\ref{eq:ef}, along with values for the electron source population number density and temperature as discussed further below, are thus used to determine the precipitating energy flux resulting from the current system discussed above. The total precipitating power for each hemisphere $P_e$ is then obtained by integration of \ensuremath{E_f} over the hemisphere, i.e.\ \begin{equation} P_e=\int_0^{90}2\pi R_{UCD}^2\sin\theta_i \;E_f \;d\,\theta_i\;\;, \label{eq:pe} \end{equation} \noindent and we finally obtain the spectral luminosity using \begin{equation} L_r=\frac{P_e}{100 \: \Delta \nu}\;\;. \label{eq:pr} \end{equation} \noindent where we assume that the beam from only one hemisphere is observable at any one time and we take the electron cyclotron maser instability to have a generation efficiency of \ensuremath{\sim}1\%, a value observed at both Jupiter and Saturn. Specifically, at Saturn this value was directly measured during the traversal of the Cassini spacecraft through the source region of the Saturn Kilometric Radiation (SKR) \citep{lamy11a}, and at Jupiter the radio power output is $\sim$1\% of the total precipitating electron power, as measured from observations of the UV aurora and computed theoretically \citep{gustin04a,clarke09a,cowley05a}. The bandwidth $\Delta \nu$ is assumed to be equal to the electron cyclotron frequency in the polar ionosphere, i.e.\ \begin{equation} \Delta\nu=\frac{eB_i}{2\pi m_e}\;\;, \label{eq:delnu} \end{equation} \noindent an approximation validated by observations of solar system planets. For example, \cite{zarka98a} argue that the jovian b-KOM, HOM and DAM emissions are a single radio component, thus with a frequency range of $\sim$10~kHz to $\sim$40~MHz, and at Saturn the SKR extends from a few kHz to $\sim$1200~kHz. This arises because the CMI occurs at all points below the field-aligned voltage drop, situated at least a few planetary radii up the field line, and the top of the ionosphere at 1 planetary radius. The emission is generated at the local electron cyclotron frequency, such that the bandwidth is then determined by the difference between the field strengths at these two locations. The strength of a dipole field drops off approximately as cube of the distance along the field line, such that at altitudes of only $\sim$1 planetary radius up the field line the frequency drops by an order of magnitude. The bandwidth is thus reasonably well represented by the high frequency cutoff, which is also the frequency used to measure the magnetic field strength in the polar region. \begin{figure} \noindent\includegraphics[width=83mm]{ucd.eps} \caption{ Figure showing profiles of model M-I coupling current system parameters. Specifically we show (a) the plasma angular velocity normalised to that of the UCD $(\omega/\Omega_{UCD})$, (b) the azimuth-integrated ionospheric Pedersen current $I_P$ in GA, (c) the field-aligned current density at the top of the ionosphere $j_{\|i}$ in $\mathrm{mA\;m^{-2}}$, (d) the minimum field-aligned voltage required to drive the current plotted in panel (c) in MV, and (e) the resulting precipitating electron energy flux $E_f$ in $\mathrm{kW\;m^{-2}}$, all plotted versus ionospheric co-latitude $\theta_i$. The Pedersen conductance employed is $\Sigma_P^*=0.5$~mho and the source electron number density and thermal energy are jovian values of 0.01~cm$^{-3}$ and 2.5~keV. The horizontal dotted lines in panels (a) and (c) indicate values of $(\omega/\Omega_{UCD})=1$ and $j_{\|i}=0$~$\mathrm{mA\;m^{-2}}$, respectively. } \label{fig:profs} \end{figure} \subsection{Angular velocity profile} We now discuss the model for the plasma angular velocity used in this paper. Following previous work \citep{cowley05a}, the gradient in the angular velocity is conveniently represented by a tanh function with respect to co-latitude, given by \begin{eqnarray} \lefteqn{\left(\frac{\omega}{\Omega_{UCD}}\right)= \left(\frac{\omega}{\Omega_{UCD}}\right)_L + }\nonumber \\ &\frac{1}{2}\left(1 - \left(\frac{\omega}{\Omega_{UCD}}\right)_L\right) \left(1+\tanh\left(\frac{\theta_i - \theta_{ic}}{\Delta\theta_i}\right)\right)\;\;, \label{eq:omega} \end{eqnarray} \noindent where $(\omega/\Omega_{UCD})_L$ is the low value to which the angular velocity transitions near the magnetic pole from rigid corotation at lower latitudes, and $\theta_{ic}$ and $\Delta\theta_i$ are the latitudinal centre and half-width of the transition region, respectively. We initially employ values for these parameters which give a Jupiter-like angular velocity profile, which on that planet generates the field-aligned currents which cause the main auroral oval and the HOM, b-KOM and non-Io-DAM radio emissions, i.e.\ $(\omega/\Omega_p)_L = 0.25$, $\theta_{ic}=15^\circ$ and $\Delta\theta_i = 0.5^\circ$. The actual form of the angular velocity gradient depends on the physical processes causing the shear, as is examined further in the Discussion, and in the results below we also examine the effects on the radio power of taking different values for these parameters. \\ \section{Results} \label{sec:results} \subsection{Results with representative values of system parameters} \label{sec:spot} In this section we present results of the model using appropriate spot values for the various model parameters as discussed above, before going on in the next section to examine ranges in the parameter space. Figure~\ref{fig:profs} shows representative results, in which we employ a uniform ionospheric field strength of 0.3~T and rotation period of 2~h, in conformity with typical values obtained from the bandwidth and modulation of the UCD radio emissions \citep{hallinan:2008aa}. First, Figure~\ref{fig:profs}a shows the angular velocity profile employed, given by equation~\ref{eq:omega}. Moving from large to small co-latitudes (or equivalently on field lines which thread the equatorial plane at increasing radial distances), the plasma angular velocity initially near-rigidly corotates, and transitions over $\sim$1$^\circ{}$ at $15^\circ{}$ co-latitude to a quarter of rigid corotation at smaller co-latitudes. The resulting equatorward azimuth-integrated ionospheric field-perpendicular (Pedersen) current given by equation~\ref{eq:totip} is shown in Figure~\ref{fig:profs}b, increasing from small values near the pole to a value of $\sim$58~GA, before falling rapidly due to the plasma angular velocity gradient shown in Figure~\ref{fig:profs}a. The resulting field-aligned current given by equation~\ref{eq:jpari} is shown in Figure~\ref{fig:profs}c, which is negative, i.e.\ downward, near the pole and switches to a significant peak of positive, i.e.\ upward, field-aligned current values, reaching amplitude $\sim$$0.7~\mathrm{mA\;m^{-2}}$, centred on $15^\circ{}$. This is the region of auroral field-aligned current that sustains the unstable electron distributions responsible for the CMI and thus the radio emissions. The required field-aligned voltage given by equation~\ref{eq:phi} is then shown in Figure~\ref{fig:profs}d, peaking at $\sim$11~MV, and the corresponding precipitating electron energy flux given by equation~\ref{eq:ef} is shown in Figure~\ref{fig:profs}e, peaking at $\sim$$9~ \mathrm{kW\;m^{-2}}$. Integrating this energy flux and converting to spectral luminosity as discussed above yields a radio power output of $\sim$$1.1~\mathrm{MW\;Hz^{-1}}$, i.e.\ consistent with the typical observed radio luminosities of UCDs. We also note that the bandwidth of the radio emission, equal to the electron cyclotron frequency of $\sim$8.4~GHz, is perforce similar to that observed, since the ionospheric field used is determined from the measured radio frequencies.\\ \begin{figure}[t] \noindent\includegraphics[width=83mm]{contours.eps} \caption{ Plots showing the spectral luminosity computed in the model for different pairs of values of the source electron population number density $n$ and thermal energy $W_{th}$. In panels (a-d) we use Pedersen conductances of $\Sigma_P^*=$~0.005, 0.05, 0.5, and 5 mho, respectively. The crosses indicate the jovian values employed in Section~\ref{sec:fav}, and the black contour indicates where $L_r=1\;\mathrm{MW\;Hz^{-1}}$, comparable to observed values. } \label{fig:contours} \end{figure} \subsection{Parameter space investigation} \label{sec:params} In the above section we have used reasonable values for a number of system parameters based on observations at Jupiter in order to demonstrate the model and indicate how an angular velocity gradient leads to radio luminosity that, for the case of fast-rotating UCDs, is comparable to that observed. However, the only parameters appropriate to this model that are known from observations, besides the radio power, are the polar magnetic field strength and rotation rate of such objects. The other parameters, such as the form of the angular velocity profile, the density and temperature of the source electron population and the ionospheric conductance are not known. Here we therefore examine ranges of these parameters to determine what values are consistent with the observed radio emission. We first examine the effects of the conductance and source electron population parameters, given the angular velocity profile given by equation~\ref{eq:omega}, and we go on to investigate the effect of changing the form of the angular velocity profile. \\ In each panel of Figure~\ref{fig:contours} we show the radio spectral luminosities obtained from the model using different pairs of values for the source electron thermal energy $W_{th}$ and number density $n$. The values adopted above obtained from Voyager observations of the plasma outside the current sheet at Jupiter, i.e.\ $~$2.5~keV and $\sim$0.01~$\mathrm{cm^{-3}}$, respectively, are shown by the crosses in each panel. The low density arises since the cold plasma is centrifugally confined to the equatorial plane, and only the hot, rarefied plasma population has a centrifugal confinement scale height large enough to populate the entire flux tube \citep{hill76a, caudal86}. At high latitudes, the low beta plasma, combined with the field-aligned voltage and loss of particles to the ionosphere, is then favourable for generation of the CMI. We note that previous estimates of the electron number density and temperature on the flux tubes producing the CMI at UCDs have been obtained, i.e.\ 1.25-5$\times10^5\;\mathrm{cm^{-3}}$ and 1-5$\times10^7$~K (equivalent to $\sim$0.86~keV) by \cite{yu:2011aa}, but these assume that the CMI is produced by an unaccelerated loss-cone population. We suggest, however, that at fast rotating UCDs, at which the magnetospheric plasma is likely to be significantly more centrifugally confined than at Jupiter, the high latitude plasma parameters may possibly be closer to the jovian case. In light of the uncertainty surrounding the source plasma population, we show the results of the model using the angular velocity profile given by equation~\ref{eq:omega}, for wide ranges of these parameters, i.e.\ 5 orders of magnitude either side of the jovian values. In panels (a-d) we employ $\Sigma_P^*=$~0.005, 0.05, 0.5, and 5 mho, respectively, reflecting the uncertainty in the Pedersen conductance. The contour labelled ``6'' indicates where the radio luminosity is equal to 1~$\mathrm{MW\;Hz^{-1}}$, representative of the observed luminosities. \\ \begin{figure} \noindent\includegraphics[width=83mm]{params.eps} \caption{ Plots showing the spectral luminosity computed in the model for different values of the angular velocity profile parameters. These are (a) the low value to which the angular velocity transitions from rigid corotation $(\omega / \Omega_p)_L$, (b) the latitudinal half-width of the transition region $\Delta\theta_i$, and (c) the centre of the transition region $\theta_{ic}$. The horizontal dotted lines are located at $L_r = 1\;\mathrm{MW\;Hz^{-1}}$. As previously, source electron population parameters are fixed at jovian values. } \label{fig:params} \end{figure} \begin{figure}[h] \noindent\includegraphics[width=83mm]{cases.eps} \caption{ Figure showing modelled (pluses) and observed (crosses) radio luminosities of the three confirmed UCD sources of CMI-generated radio emissions versus each UCD's angular velocity. Model parameters are given in Table~\ref{tab:ucds} } \label{fig:cases} \end{figure} It is apparent that over the majority of the parameter space plotted, the luminosity increases with decreasing $n$ and increasing $W_{th}$, with a change of slope as the relativistic terms dominate above $W_{th}\simeq511$~keV. Where the slope reverses in the top right corner of each panel the field-aligned current can be carried by an unaccelerated population. For simplicity, in this region we have assumed that in the auroral region the loss-cone is filled by strong pitch angle diffusion, and that the required net current is then created by a corresponding upward flux of electrons. We have also taken the same CMI generation efficiency of 1\% as for the accelerated population, despite the less favourable loss-cone distribution in this case. Both these assumptions will likely lead to overestimates of the radio flux from this region. We also note that the previous estimates of the source population parameter values discussed above \citep{yu:2011aa} are off the top of each panel, thus clearly in the unaccelerated regime. In any case, it is apparent in the accelerated region that a wide range of parameter values are able to produce the required radio luminosity, in general requiring hotter, less dense plasma for lower Pedersen conductance. The plot also confirms the results shown above, i.e.\ that typically jovian values for the source plasma population parameters and the Pedersen conductance lead to observed UCD radio luminosities. \\ We now consider the effect of the form of the angular velocity profile, and show in Figure~\ref{fig:params} the radio luminosities obtained when the parameters in equation~\ref{eq:omega} are varied individually. Specifically, from top to bottom the parameters varied are (a) the low value to which the angular velocity transitions from rigid corotation $(\omega / \Omega_p)_L$, (b) the latitudinal half-width of the transition region $\Delta\theta_i$, and (c) the centre of the transition region $\theta_{ic}$. As a guide, the horizontal dotted lines highlight $1\;\mathrm{MW\;Hz^{-1}}$, which delimits the parameter ranges reasonably consistent with observations. Thus, an angular velocity profile which transitions to $\lesssim70$\% of rigid corotation across a region of half-width $\lesssim2.5^\circ$ at a co-latitude $\gtrsim9^\circ$ will produce radio powers consistent with observations. These are not particularly tight constraints, indicating that the model does not require fine tuning to produce the required emissions. Future modelling of the radio beams, along the lines of previous work that has concentrated on localised active regions \citep{kuznetsov:2012aa}, from angular velocity transition regions of different geometries would be useful to constrain the angular velocity profiles further.\\ \begin{table*} \begin{center} \caption{Properties of the UCDs considered in this paper.\label{tab:ucds}} \begin{tabular}{lccc} \tableline\tableline Property & TVLM 513 & LSR J1835 & 2M J0036 \\ \tableline Radius $R_{UCD}$ / $\mathrm{R_J}$ \tablenotemark{a}& 1.005 & 1.142 & 0.927 \\ Rotation period $\tau$ / h \tablenotemark{a}& 1.96 & 2.84 & 3.08 \\ Polar ionospheric field strength $B_i$ / T \tablenotemark{b}& 0.3 & 0.3 & 0.17 \\ Distance $d$ / pc \tablenotemark{a} & 8.8 & 5.7 & 10.6 \\ Quiescent spectral luminosity $L_{r\:q}$ / $\mathrm{MW\;Hz^{-1}}$ \tablenotemark{c}& 5.77 & 2.77 & 1.23 \\ Mean peak spectral flux density $F_r$ / mJy & 4.2\tablenotemark{d}& 2.4\tablenotemark{e} & 0.5\tablenotemark{f}\\ Pulse duty cycle $\Delta \tau / \tau$ & 0.05\tablenotemark{g}& 0.1\tablenotemark{e} & 0.3\tablenotemark{f}\\ Estimated peak spectral luminosity $L_{r}$ / $\mathrm{MW\;Hz^{-1}}$ \tablenotemark{h} & 2.6 & 0.8 & 1.3 \\ Pedersen conductance $\Sigma_P^*$ / mho \tablenotemark{i} & 0.5 & 0.3 & 0.5 \\ \tableline \end{tabular} \tablenotetext{a}{From Table~1 of \cite{hallinan:2008aa}} \tablenotetext{b}{From \cite{hallinan:2008aa}} \tablenotetext{c}{From Table~2 of \cite{hallinan:2008aa}} \tablenotetext{d}{From Figure~1 of \cite{hallinan:2007aa}} \tablenotetext{e}{From Figure~1 of \cite{hallinan:2008aa}} \tablenotetext{f}{From Figure~2 of \cite{hallinan:2008aa}} \tablenotetext{g}{From Figure~3 of \cite{hallinan:2007aa}} \tablenotetext{h}{From equation~\ref{eq:lr}} \tablenotetext{i}{Employed in Figure~\ref{fig:cases}} \end{center} \end{table*} \subsection{Comparison with observations} We now compare the model spectral luminosities with the confirmed sources of CMI-induced radio emissions discussed by \cite{hallinan:2008aa}, i.e.\ TVLM 513-46546, LSR J1835+3259, and 2MASS J00361617+1821104 (hereafter TVLM 513, LSR J1835 and 2M J0036), which have relevant properties listed in Table~\ref{tab:ucds}. We estimate the spectral luminosities of the CMI-induced radio emission from the peak amplitudes and widths plotted by \cite{hallinan:2007aa, hallinan:2008aa}. As discussed above, CMI-induced radio emission is strongly beamed, into 1.6~sr for the case of Jupiter's HOM and DAM emissions, resulting in a duty cycle of $\sim$0.14 for these emissions. We estimate the beam width $\sigma$ for these UCDs by simply scaling to the jovian emissions via \begin{equation} \sigma = \left(\frac{\Delta\tau / \tau}{0.14}\right) 1.6\;\;, \label{eq:beam} \end{equation} \noindent where the pulse duty cycle $\Delta\tau / \tau$ is given in Table~\ref{tab:ucds}. This, however, does assume that the angular velocity profile, and thus that of the field-aligned currents, are the same for each dwarf, which may not be the case. The spectral luminosity then follows from the peak flux density $F$ via \begin{equation} L_r = F\sigma d^2\;\;, \label{eq:lr} \end{equation} \noindent where $d$ is the distance to the object from the Earth. This is shown by the crosses in Figure~\ref{fig:cases}. In comparison, the modelled spectral luminosity, shown by the pluses, is computed using the radii, rotation periods, polar ionospheric field strengths and Pedersen conductances as listed in Table~\ref{tab:ucds}. The Pedersen conductances employed for TVLM 513 and 2M J0036 are 0.5~mho, as is also used in Figure~\ref{fig:profs}, but it was found that 0.3~mho yields better agreement for LSR J1835. It is also worth noting that employing 0.8~mho for all three dwarfs yields spectral luminosities consistent to within $\sim7$\% of the quiescent values (computed assuming isotropic emission) given by \cite{hallinan:2008aa}. \\ We also note the potential CMI-induced emissions from the L dwarf binary 2MASSW J0746425+ 200032 \citep{berger:2009aa} and the T6.5 brown dwarf 2MASS J10475385+2124234 \citep{route:2012aa}. For simplicity we do not consider the dwarf binary since the behaviour of such a system could be significantly more complex than is represented by our simple axisymmetric model. Regarding the latter dwarf, two highly polarised bursts were observed at 4.75~GHz (implying a field strength of $\sim$0.17~T), with spectral flux densities of 2.7~$\pm$~0.2~mJy (where 1 Jansky = $10^{-26}\;\mathrm{W\;m^{-2}\;Hz^{-1}}$) and 1.3~$\pm$~0.2 mJy, respectively. Assuming the radio emission is beamed into 1.6~sr and noting the object's 10.3~pc distance, these flux densities imply a mean spectral luminosity of $(3.3 \pm 0.7)~\mathrm{MW\;Hz^{-1}}$. This observation is not included in Figure 2 since no rotation period is known, although it is likely to be at least 2~h since only one burst was observed in each $\sim$2~h observing interval. Using the parameter values employed in this paper, our model indeed requires rotation periods of 2.1-2.8~h to produce radio powers consistent with the above luminosities, such that further observations of this object are required to test this scenario. \\ \section{Summary and Discussion} In this paper we have developed an axisymmetric model of the M-I coupling currents arising from departure from rigid corotation in the magnetospheres of UCDs, based on that used previously to study Jupiter's auroral oval. This simple model generates radio emissions whose power, bandwidth, modulation period and duty cycle are consistent with those observed. The implication is that these UCDs are coupling with their space environments via magnetic fields, and that this coupling reduces the angular velocity of the high latitude ionopsheric regions, thus generating intense auroral emissions including radio emissions. \\ In this study the angular velocity profile is used as the input to the model, whereas in reality the form of the profile will be dictated by the physical processes governing the magnetospheric flows. The magnetospheres of solar system planets are driven by two sources of momentum, i.e.\ the solar wind and planetary rotation. Which of these is dominant depends on the individual circumstances, with Earth's magnetosphere being dominated by the solar wind, while Jupiter's and Saturn's are rotationally dominated. Previous work \citep{kuznetsov:2012aa} has assumed that the radio from UCDs is emitted from field lines mapping to $\rho_e = 2R_{UCD}$, where $\rho_e$ represents the equatorial radial distance from the magnetic axis, on the basis that at this distance the gravitational energy density equals the corotational energy density, such that it is hypothesised that beyond this distance any initially corotating plasma becomes centrifugally unstable and flows approximately radially away from the object \citep{ravi:2011aa}. It is worth noting, however, that at Jupiter this distance is also $\sim$2~$\mathrm{R_J}$, and yet the dominant plasma flow is azimuthal out to the magnetopause at $\sim$40--100~$\mathrm{R_J}$ and no auroral emission is generated on field lines mapping to $\rho_e\simeq2\;\mathrm{R_J}$. \\ A more complete equation of motion is \begin{equation} \rho \left(\frac{d\mathbf{v}}{dt} - \mathbf{g}\right) +\nabla p = \mathbf{j} \times \mathbf{B}\;\;, \label{eq:motn} \end{equation} \noindent where $\rho$ is the plasma mass density, $d\mathbf{v}/dt$ is the acceleration of the plasma bulk flow with respect to inertial space, $\mathbf{g}$ is the acceleration due to gravity, $p$ is the plasma pressure, $\mathbf{j}$ is the current density and $\mathbf{B}$ the magnetic field, and we note that in the corotating frame the $d\mathbf{v}/dt$ term comprises the convective and local accelerations, centrifugal, Coriolis and differential rotation effects \citep{vasyliunas83}. The net result of the above force balance is that Jupiter's magnetosphere is radially distended into a magnetodisc structure, which doubles its size from that which might be expected from the planetary dipole alone \citep{caudal86,nichols11b}. Plasma in Jupiter's magnetosphere is, however, centrifugally unstable and diffuses radially outward via flux tube interchange motions \citep{bespalov06} before being lost down the tail via the pinching off of plasmoids at around $\sim$100~$\mathrm{R_J}$ \citep{vogt10a}. If there were no ionospheric torque on the magnetospheric plasma its angular velocity would decrease with $1/\rho^2$, where $\rho$ is the distance from the spin axis, but the finite ionospheric Pedersen conductance allows a large-scale M-I coupling current system such as that discussed in this paper to flow, the $\mathbf{j} \times \mathbf{B}$ force of which balances the neutral drag in the ionosphere and maintains partial corotation throughout the closed magnetosphere. The equation of motion which describes the angular velocity of the plasma under steady outflow from an internal source was derived originally by \cite{hill79}, and is given by \begin{eqnarray} \lefteqn{\frac{\rho_e}{2}\frac{d}{d\rho_e}\left(\frac{\omega}{\Omega_{UCD}}\right)+\left(\frac{\omega}{\Omega_{UCD}}\right)=}\nonumber \\ &\frac{4\pi \Sigma_P^*F_e|B_{ze}|}{\dot{M}}\left(1-\frac{\omega}{\Omega_{UCD}}\right)\;\;, \label{eq:hp} \end{eqnarray} \noindent where $F_e$ is the equatorial value of the the poloidal flux function $F$, related to the magnetic field $\mathbf{B}$ by $\mathbf{B}=(1/\rho)\nabla F \times \hat{\varphi}$, $|B_{ze}|$ is the magnitude of the north-south magnetic field threading the equatorial plane, and $\dot{M}$ is the plasma mass outflow rate. Analytic solutions to equation~\ref{eq:hp} can be obtained, depending on the form of the magnetic field, along with characteristic distances over which the plasma breaks from rigid corotation \citep{nichols03}, called the `Hill distance' (not to be confused with the radius of the Hill sphere within which a body's gravitational force dominates). In the absence of knowledge of the magnetodisc structure of these UCDs, taking the dipole case represents a lowest order approximation, for which the Hill distance is given by \begin{equation} \left(\frac{\rho_{H}}{R_{UCD}}\right) = \left(\frac{2\pi\Sigma_P^*B_{eq}^2 R_{UCD}^2}{\dot{M}}\right)^{1/4}\;\;, \label{eq:rhoh} \end{equation} \noindent where $B_{eq}$ is the equatorial field at the surface of the UCD. Employing $\Sigma_P^*=0.5$~mho and the canonical value of $\dot{M} = 1000\;\mathrm{kg\;s^{-1}}$ for the Io source rate, we then have $\rho_H \simeq775\;\mathrm{R_{UCD}}$. \\ In comparison, the size of the magnetosphere $R_{mp}$ is estimated by considering pressure balance between the magnetic field pressure of the compressed planetary dipole just inside the magnetopause and the magnetic, thermal, and dynamic pressures of the interstellar medium on the outside, i.e.\ \begin{equation} \left(\frac{\ensuremath{R_{mp}}}{R_{UCD}}\right)=\left(\frac{k_m^2B_{eq}^2}{2\mu_\circ (p_{th} + p_{dyn} + p_{B})}\right)^{1/6}\;\;, \label{eq:rmp} \end{equation} \noindent where $k_m$ is the factor by which the magnetopause field is enhanced by magnetopause currents, given by $\ensuremath{\sim}2.44$ for a frontal boundary of realistic shape, i.e.\ lying between the values of 2 and 3 appropriate to planar and spherical boundaries, respectively \citep{mead64a,alexeev05a}. Taking plasma number density, temperature and magnetic field strength values for both the local interstellar cloud and the local bubble, along with a representative velocity for the UCD of 50~$\mathrm{km\;s^{-1}}$ with respect to the local medium \citep{vanhamaki:2011aa}, yields values of $R_{mp} \simeq 713$ and $820~\mathrm{R_{UCD}}$ for the local interstellar cloud and local bubble, respectively. Thus, on this basis corotation breakdown may only be marginal inside the closed magnetosphere, although a more detailed analysis of the force balance should be undertaken in the future to determine whether this would be the case for a realistic current sheet field structure. \\ It is also worth mentioning that while at least TVLM 513 may possibly have a companion located at $20-50\;\mathrm{R_{UCD}}$ \citep{forbrich:2009aa}, evidence for satellite-induced radio emission has not been found \citep{kuznetsov:2012aa}, but in this regard we note that at Jupiter the brightest and most significant auroral form resulting from the Io plasma source is the main oval, whose apparent modulation is at the planetary period, not Io's orbital period. If near-rigid corotation is then maintained throughout the closed magnetosphere, an interesting point is that the linear velocities of the plasma near the magnetopause may reach a significant fraction of the speed of light, i.e.\ $\sim$$0.17c$ for rigid corotation with a period of 2~h at $800\;\mathrm{R_{UCD}}$. An order of magnitude approximation of the limiting distance at which plasma content can be maintained in corotation by inward magnetic stress is where the Alfv\'en speed equals the corotation speed. The former is, of course, unknown since we do not have knowledge of the plasma density, but at high latitudes in Jupiter's magnetosphere it can approach the speed of light. If the limiting distance does lie significantly within the magnetopause then a super-Alfv\'enic radially-outward planetary wind may indeed form, at which point the distinction between open and closed field lines could become rather blurred \citep{kennel77}, although it is worth noting that at Jupiter this distance is also within the magnetosphere and no swift radial outflow occurs at this location. \\ The amount of flux `truly' open to the interstellar medium will depend on the reconnection rates on the leading and trailing sides of the flow with respect to the surrounding medium. Theories of magnetic reconnection have not yet reached the point where ab initio calculations of the rate of flux transport can be made, however it is known that reconnection occurs where a significant magnetic shear is present. Without in situ measurements we do not of course know what the interstellar magnetic field is at the magnetopause, but it is likely to be piled up against the boundary, since the $\sim$50 $\mathrm{km\;s^{-1}}$ proper motions typical of low mass stars are similar to the velocities observed in the shocked solar wind plasma in Jupiter's magnetosheath just upstream of the magnetopause \citep[e.g. ][]{siscoe:1980aa}. At planets in the solar system the open-closed field line boundaries typically lie between $\sim$10--15$^\circ$ co-latitude. The angular velocity of the open field line region \citep{isbell84} is given by \begin{equation} \left(\frac{\omega}{\Omega_{UCD}}\right)_{Open} = \frac{\mu_0 \Sigma_P^*v_{LIM}}{1 + \mu_0\Sigma_P^* v_{LIM}}\;\;, \label{eq:isbell} \end{equation} \noindent where $v_{LIM}$ is the velocity of the UCD through the local interstellar medium, which for the above parameter values yields $(\omega/\Omega_{UCD})_{Open}$ = 0.05, and may be much lower in the polar region if there is little aurorally-generated conductivity. Therefore, an open-closed field line boundary is a possible cause of the angular velocity gradient discussed in this paper. We note that \cite{schrijver:2009aa} considered the interaction of the magnetosphere with the interstellar medium, and ruled it out, by showing that the kinetic energy flux of the impinging medium is much less than typical X-ray and H-$\alpha$ luminosities. However, this does not consider the rotation of the body, which is a crucial property of these UCDs as discussed above. Here we postulate that the interaction may instead be such that magnetic reconnection occurs on the front side of the magnetosphere and creates an essentially stagnant region of open field lines, whose angular velocities are given by equation~\ref{eq:isbell}, and the transition from corotating closed flux to sub-rotating open field lines drives field-aligned currents. This is the current system described for Saturn's auroras by e.g. \cite{cowley03d} and \cite{cowley04a} and possibly a component of Jupiter's auroras poleward of the main oval \citep{cowley03f,cowley05a,cowley07}. Finally, it is worth noting that the auroral and radio emissions in our solar system caused by the above-discussed processes are all highly variable in terms of power, morphology and, in Saturn's case, modulation period, due to a number of reasons including varying interplanetary medium parameters and internal plasma source rates, such that it may be expected that the radio emissions from UCDs are similarly variable. \acknowledgments JDN was supported by an STFC Advanced Fellowship. SWHC was supported by STFC grant ST/H002480/1. JDN wishes to thank G. Hallinan for helpful discussions regarding the radio observations of UCDs.
\section{Introduction}\label{sec:intro} In the past few decades, a variety of wavelets that provide a complete and stable multiscale representation of $L_{2}(\mathbb{R}^d)$ have been developed. The wavelet decomposition is very efficient from a computational point of view, due to the fast filtering algorithm. A fundamental property of traditional wavelet basis functions is that they behave like multiscale derivatives \cite{mallat09,meyer92}. Our purpose in this paper is to extend this concept by constructing wavelets that behave like a given Fourier multiplier operator $\Lop$, which can be more general than a pure derivative. In our approach, the multiresolution spaces are characterized by generalized B-splines associated with the operator, and we show that, in a certain sense, the wavelet inherits properties of the operator. Importantly, the operator-like wavelet can be constructed directly from the operator, bypassing the scaling function space. What makes the approach even more attractive is that, at each scale, the wavelet space is generated by the shifts of a single function. Our work provides a generalization of some known constructions including: cardinal spline wavelets \cite{chui91}, elliptic wavelets \cite{micchelli91}, polyharmonic spline wavelets \cite{vandeville05,vandeville10}, Wirtinger-Laplace operator-like wavelets \cite{vandeville08}, and exponential-spline wavelets \cite{khalidov06}. In applications, it has been observed that many signals are well represented by a relatively small number of wavelet coefficients. Interestingly, the model that motivates our wavelet construction explains the origin of this sparsity. The context is that of sparse stochastic processes, which are defined by a stochastic differential equation driven by a (non-Gaussian) white noise. Explicitly, the model states that $\Lop s=w$ where the signal $s$ is a tempered distribution, $\Lop$ is a shift-invariant Fourier multiplier operator, and $w$ is a sparse white noise \cite{unser11sm}. The wavelets we construct are designed to act like the operator $\Lop$ so that the wavelet coefficients are determined by a generalized B-spline analysis of $w$. In particular, we define an interpolating spline $\phi$, corresponding to $\Lop^*\Lop$, from which we derive the wavelets $\psi=\Lop^*\phi$. Then the wavelet coefficients are formally computed by the $L_2$ inner product \begin{equation*} \langle s,\psi \rangle = \langle s,\Lop^* \phi \rangle = \langle \Lop s, \phi \rangle = \langle w,\phi \rangle . \end{equation*} Sparsity of $w$ combined with localization of the interpolating spline $\phi$ results in sparse wavelet coefficients \cite{unser_part1}. This model is relevant in medical imaging applications, where good performance has been observed in approximating functional magnetic resonance imaging and positron emission tomography data using operator-like wavelets that are tuned to the hemodynamic or pharmacokinetic response of the system \cite{khalidov11,verhaeghe08}. Our construction falls under the general setting of pre-wavelets, which are comprehensively covered by de Boor, DeVore, and Ron in \cite{deboor93}. Two distinguishing properties of our approach are its operator-based nature and the fact that it is non-stationary. Related constructions have been developed for wavelets based on radial basis functions \cite{chui96aw,chui96wa,stoeckler93}. In fact, \cite{chui96wa} also takes an operator approach; however, the authors were focused on wavelets defined on arbitrarily spaced points. This paper is organized as follows. In Section \ref{sec:prelim}, we formally define the class of admissible operators and the lattices on which our wavelets are defined. In Section \ref{sec:multires}, we construct the non-stationary multiresolution analysis (MRA) that corresponds to a given operator $\Lop$ and derive approximation rates for functions lying in Sobolev-type spaces. Then, in Section \ref{sec:wave}, we introduce the operator-like wavelets and study their properties; in particular, we derive conditions on $\Lop$ that guarantee that our choice of wavelet yields a stable basis at each scale. Under an additional constraint on $\Lop$, we use this result to define Riesz bases of $L_2(\mathbb{R}^d)$. In Section \ref{sec:decor}, we prove a decorrelation property for families of related wavelets. In Section \ref{sec:conclusion}, we present connections to prior constructions, and we conclude with some examples of operator-like wavelets. \section{Preliminaries}\label{sec:prelim} The primary objects of study in this paper are Fourier multiplier operators and their derived wavelets. The operators that we consider are shift-invariant operators $\Lop$ that act on $L_2(\mathbb{R}^d)$, the class of square integrable functions $f:\mathbb{R}^d \to \mathbb{C}$. The action of such a Fourier multiplier operator is defined by its symbol $\widehat{L}$ in the Fourier domain, with \begin{equation*} \Lop f = \left( \widehat{L} \widehat{f} \right)^\vee. \end{equation*} The symbol $\widehat{L}$ is assumed to be a measurable function. The adjoint of $\Lop$ is denoted as $\Lop^*$, and its symbol is the complex conjugate of $\widehat{L}$; i.e., the symbol of $\Lop^*$ is $\widehat{L}^*$. In the previous equation, we used $\widehat{f}$ to denote the Fourier transform of $f$ \begin{equation*} \widehat{f}(\bm{\omega}) = \int_{\mathbb{R}^d} f(\bm{x}) e^{-i\bm{x}\cdot\bm{\omega}} {\rm d}\bm{x}. \end{equation*} We use $g^\vee$ to denote the inverse Fourier transform of $g$. Pointwise values of $\widehat{L}$ are required for some of our analysis, so we restrict the class of symbols by requiring continuity almost everywhere. Additionally, we would like to have a well-defined inverse of the symbol, so $\widehat{L}$ should not be zero on a set of positive measure. To be precise, we define the class of admissible operators as follows. \begin{definition}\label{def:admis} Let $\Lop$ be a Fourier multiplier operator. Then $\Lop$ is admissible if its symbol $\widehat{L}$ is of the form $f/g$, where $f$ and $g$ are continuous functions satisfying: \begin{enumerate} \item The set of zeros of $fg$ has Lebesgue measure zero; \item The zero sets of $f$ and $g$ are disjoint. \end{enumerate} \end{definition} Notice that each such operator defines a subspace of $L_2$, consisting of functions whose derivatives are also square integrable, and our approximation results focus on these spaces. \begin{definition} An admissible operator $\Lop$ defines a Sobolev-type subspace of $L_2(\mathbb{R}^d)$: \begin{equation*} W_2^\Lop(\mathbb{R}^d) := \left\{f\in L_2({\mathbb R^d}):\norm{f}_{W_2^{\Lop}}<\infty\right\}, \end{equation*} where \begin{equation*} \norm{f}_{W_2^{\Lop}} := \left( \int_{\mathbb R^d}\abs{\widehat f(\bm{\omega})}^2 \left(1+\abs{\widehat L(\bm{\omega})}^2\right)\rmd\bm{\omega} \right)^{1/2}. \end{equation*} \end{definition} Having defined the class of admissible operators, we must consider the lattices on which the multiresolution spaces will be defined. It is important to use lattices which are nested, so we consider those defined by an expansive integer matrix. Specifically, an integer matrix $\mathbf{A}$, whose eigenvalues are all larger than 1 in absolute value, defines a sequence of lattices \begin{equation*} \mathbf{A}^j\mathbb{Z}^d = \{\mathbf{A}^j\bm{k}:\bm{k}\in \mathbb{Z}^d \} \end{equation*} indexed by an integer $j$. Using \cite{kalker99} as a reference, we recall some results about lattices generated by a dilation matrix. First, we know that $\mathbf{A}^{j}\mathbb{Z}^d$ can be decomposed into a finite union of disjoint copies of $\mathbf{A}^{j+1}\mathbb{Z}^d$; there are $\abs{\det(\mathbf{A})}$ vectors $\{\bm{e}_{l}\}_{l=0}^{\abs{\det(\mathbf{A})}-1}$ such that \begin{equation*} \bigcup_l \left(\mathbf{A}^j\bm{e}_{l} + \mathbf{A}^{j+1}\mathbb{Z}^d\right) = \mathbf{A}^{j} \mathbb{Z}^d, \end{equation*} and using this notation, our convention will be to set $\bm{e}_{0}=\bm{0}$. There are also several important properties that arise when using Fourier techniques on more general lattices. A lattice in the spatial domain corresponds to a dual lattice in the Fourier domain, and the dual lattice of $\mathbf{A}^j\mathbb{Z}^d$ is given by $2\pi(\mathbf{A}^T)^{-j}\mathbb{Z}^d$. Also relevant is the notion of a fundamental domain, which for $\mathbf{A}^j\mathbb{Z}^d$ is a bounded, measurable set $\Omega_j$ satisfying \begin{equation*} \sum_{\bm{k} \in \mathbb{Z}^d} \chi_{\Omega_j}(\bm{x}+\mathbf{A}^j\bm{k}) = 1 \end{equation*} for all $\bm{x}$. In this paper, we restrict our attention to lattices derived from matrices that are constant multiples of orthogonal matrices; i.e., we assume a scaling matrix $\mathbf{A}$ satisfies $\mathbf{A}=a\mathbf{R}$ for some orthogonal matrix $\mathbf{R}$ and constant $a>1$. The lattices generated by these matrices have some additional nice properties. For example, the lattices generated by $\mathbf{A}$ and $\mathbf{A}^T$ are the same, and the lattices $\mathbf{A}^j\mathbb{Z}^d$ scale uniformly in every direction for $j\in \mathbb{Z}$. Also, for such matrices, there are only finitely many possible lattices generated by powers of $\mathbf{A}$; i.e., there always exists a positive integer $n$ for which $\mathbf{A}^n=a^n\mathbf{I}$. In addition to the standard dilation matrices $a\mathbf{I}$ (where $a=2,3,\dots$), there are other matrices satisfying the restriction described above. For example in two dimensions, the quincunx matrix \begin{equation*} \mathbf{A} = \left( \begin{matrix} 1&1\\ 1&-1 \end{matrix} \right) \end{equation*} is valid, and in three dimensions, one could use \begin{equation*} \mathbf{A} = \left( \begin{matrix} 2&2&-1\\ 2&-1&2\\ -1&2&2 \end{matrix} \right). \end{equation*} \section{Multiresolution Analysis}\label{sec:multires} The multiresolution framework for wavelet construction was presented by Mallat in the late 1980s \cite{mallat89}. In the following years, the notion of pre-wavelets was developed, and a more general notion of multiresolution was adopted. We consider this more general setting in order to allow for a wider variety of admissible operators. \begin{definition}\label{def:nonstat_mra} A sequence $\{V_j\}_{j\in \mathbb{Z}}$ of closed linear subspaces of $L_2(\mathbb{R}^d)$ forms a non-stationary multiresolution analysis if \begin{enumerate} \item $V_{j+1} \subseteq V_j$; \item $\bigcup_{j\in \mathbb{Z}} V_j$ is dense in $L_2(\mathbb{R}^d)$ and $\bigcap_{j\in \mathbb{Z}} V_j$ is at most one-dimensional; \item $f\in V_j$ if and only if $f(\cdot-\mathbf{A}^j\bm{k})\in V_j$ for all $j\in \mathbb{Z}$ and $\bm{k}\in \mathbb{Z}^d$, where $\mathbf{A}$ is an expansive integer matrix; \item For each $j\in \mathbb{Z}$, there is an element $\varphi_j\in V_j$ such that the collection of translates $\{\varphi_j(\cdot-\mathbf{A}^j\bm{k}): \bm{k}\in \mathbb{Z}^d \}$ is a Riesz basis of $V_j$, i.e., there are constants $0<A_j\leq B_j<\infty$ such that \begin{equation*} A_j \norm{c}_{\ell_2}^2 \leq \norm{\sum_{\bm{k}\in \mathbb{Z}^d} c[\bm{k}] \varphi_j(\cdot-\mathbf{A}^j\bm{k}) }_{L_2(\mathbb{R}^d)}^2 \leq B_j \norm{c}_{\ell_2}^2. \end{equation*} \end{enumerate} \end{definition} Let us point out here a few remarks concerning this definition. First of all, note that we have defined our multiresolution spaces $V_j$ to be `growing' as $j$ approaches $-\infty$. Also, in the second condition we do not require the intersection of the spaces $V_j$ to be $\{0\}$. Instead, we allow it to be one-dimensional. This happens, for example, when every space is generated by the dilations of a single function; i.e., there is a $\varphi\in L_{2}(\mathbb{R}^d)$ such that \begin{equation*} V_j = \left\{\sum_{\bm{k}\in \mathbb{Z}^d} c[\bm{k}]\varphi(\cdot-\mathbf{A}^j \bm{k}): c\in \ell_2(\mathbb{Z}^d)\right\} \end{equation*} for every $j$. In order to produce non-stationary MRAs, we require additional properties on an admissible operator. Together with a dilation matrix, the operator should admit generalized B-splines (generators of the multiresolution spaces $V_j$) that satisfy decay and stability properties. As motivation for our definition, let us consider the one-dimensional example where $\Lop$ is defined by \begin{equation*} \Lop f (t) = \frac{\rmd f}{\rmd t}(t)-\alpha f(t), \end{equation*} for some $\alpha>0$. A Green's function for $\Lop$ is $\rho(t)=e^{\alpha t}H(t)$, where $H$ is the Heaviside function. In order to produce Riesz bases for the scaling matrix $\mathbf{A}=(2)$, we introduce the localization operators $\Lop_{\rmd,j}$ defined by $\Lop_{\rmd,j}f=f-e^{2^j \alpha}f(\cdot-2^j)$. Then for any $j\in\mathbb{Z}$, the exponential B-spline $\varphi_j:=\Lop_{\rmd,j}\rho$ is a compactly supported function whose shifts $\{\varphi_j(\cdot-2^jk)\}_{k\in\mathbb{Z}}$ form a Riesz basis. In the Fourier domain, a formula for $\varphi_j$ is $\widehat{L}(\omega)^{-1}\widehat{L}_{\rmd,j}(\omega)$, which is \begin{equation*} \widehat{\varphi}_j(\omega) = \frac{1-e^{2^j(\alpha-i\omega)}}{i\omega-\alpha}. \end{equation*} In this form we verify the equivalent Riesz basis condition \begin{equation*} 0 < A_j \leq \sum_{k\in \mathbb{Z}} \abs{\widehat{\varphi}_j(\cdot-2\pi 2^{-j}k) }^2 \leq B_j < \infty. \end{equation*} In fact, based on the symbol of $\Lop$, we could have worked entirely in the Fourier domain to determine appropriate periodic functions $\widehat{L}_{\rmd,j}$. With this example in mind, we make the following definition. \begin{definition}\label{def:spadm} We say that an operator $\Lop$ and an integer matrix $\mathbf{D}$ are a spline-admissible pair of order $r>d/2$ if the following conditions are satisfied: \begin{enumerate} \item $\Lop$ is an admissible Fourier multiplier operator; \item $\mathbf{D}=a\mathbf{R}$ with $\mathbf{R}$ an orthogonal matrix and $a>1$; \item There is a constant $C_{\Lop}>0$ such that \begin{equation*} C_{\Lop} \left( 1+\abs{\widehat{L}(\bm{\omega})}^2 \right) \geq \abs{\bm{\omega}}^{2r}; \end{equation*} \item For every $j\in \mathbb{Z}$, there exists a periodic function $\widehat{L}_{ \mathrm{d},j}$ such that $\widehat{\varphi}_j(\bm{\omega}):=\widehat{L}_{\mathrm{d},j}(\bm{\omega}) \widehat{L}(\bm{\omega})^{-1}$ satisfies the Riesz basis condition \begin{equation*} 0 < A_j \leq \sum_{\bm{k}\in\mathbb{Z}^d}\abs{\widehat\varphi_j(\bm{\omega}+2\pi (\mathbf{D}^T)^{-j}\bm{k})}^2 \leq B_j < \infty, \end{equation*} for some $A_j$ and $B_j$ in $\mathbb{R}$. Here, we require the periodic functions $\widehat{L}_{ \mathrm{d},j}$ to be of the form $\sum_{\bm{k} \in \mathbb{Z}^d} p_j[\bm{k}]e^{i\bm{\omega}\cdot \mathbf{D}^j \bm{k}}$ for some $p\in\ell_1(\mathbb Z^d)$. \end{enumerate} \end{definition} \begin{definition} Let $\Lop$ and $\mathbf{D}$ be a spline admissible pair. The functions $\widehat{\varphi}_j$ from Condition 4 of Definition \ref{def:spadm} are in $L_2(\mathbb{R}^d)$, and we refer to the functions \begin{equation*} \varphi_j := (\widehat{\varphi}_j)^\vee \end{equation*} as generalized B-splines for $\Lop$. \end{definition} \begin{proposition} Given a spline-admissible pair $\Lop$ and $\mathbf{D}$, the spaces \begin{equation*} V_j = \left\{\sum_{\bm{k}\in \mathbb{Z}^d} c[\bm{k}]\varphi_j(\cdot-\mathbf{D}^j \bm{k}): c\in \ell_2(\mathbb{Z}^d)\right\} \end{equation*} form a non-stationary MRA. \end{proposition} \begin{proof} The first property of Definition \ref{def:nonstat_mra} is verified using the definition of $\mathbf{D}$ and the Riesz basis conditions on the generalized B-splines $\varphi_j$. Density in $L_2(\mathbb{R}^d)$ is a result of the admissibility of $\Lop$, the Riesz basis condition on $\varphi_j$, and the inclusion relation $V_{j+1} \subseteq V_{j}$, cf. \cite[Theorem 4.3]{deboor93}. Also, there is an integer $n$ for which $\mathbf{D}^n=a^n\mathbf{I}$, and the intersection of the spaces $V_{jn}$ is at most one-dimensional by Theorem 4.9 of \cite{deboor93}. Property 3 follows from the definition of the spaces $V_j$, and lastly, Property 4 of Definition \ref{def:nonstat_mra} is valid due to Property 4 of Definition \ref{def:spadm}. \end{proof} The primary difficulty in proving spline-admissibility is verifying Condition 4, which concerns the existence of generalized B-splines. This problem is closely related to the localization (or `preconditioning') of radial basis functions for the construction of cardinal interpolants \cite{chui92}. As in that paper, the idea is to construct periodic functions $\widehat{L}_{ \mathrm{d},j}$ that cancel the singularities of $\widehat{L}^{-1}$. In one dimension, we can verify spline-admissibility for any constant-coefficient differential operator. In higher dimensions, spline admissibility holds for the Mat\'ern operators, characterized by $\widehat{L}(\bm{\omega})=(1+\abs{\bm{\omega}}^2)^{\nu/2}$, as they require no localization. In Section \ref{sec:conclusion}, we provide a less obvious example and show how this Riesz basis property can be verified. As a final point, note that if one is only interested in analyzing fine-scale spaces, Condition 4 need only be satisfied for $j$ smaller than a fixed integer $j_0$, but in this case, it is necessary to include the space $V_{j_0-1}$ in the wavelet decomposition. We close this section by determining approximation rates for the multiresolution spaces $\{V_j\}_{j\in \mathbb{Z}}$, in terms of the operator $\Lop$ and the density of the lattices generated by $\mathbf{D}^j$. In order to state this result, we define the spline interpolants for the operator $\Lop^*\Lop$, whose symbol is $\absf{\widehat{L}}^2$. The spline admissibility of this operator is the subject of the next proposition. \begin{proposition} If $\Lop$ is spline admissible of order $r>d/2$, then $\Lop^*\Lop$ is spline admissible of order $2r>d$. \end{proposition} \begin{proof} Let $\Lop$ be a spline admissible operator of order $r>d/2$. First notice that $\Lop^*\Lop$ is an admissible Fourier multiplier operator. Also, we see that $\Lop^*\Lop$ satisfies Condition 3 of Definition \ref{def:spadm} with $r$ replaced by $2r$. Therefore spline admissibility follows if we can exhibit generalized B-splines for $\Lop^*\Lop$ that satisfy the Riesz basis condition, where the integer dilation matrix is the same as for $\Lop$. To that end, let $\varphi_j$ be a generalized B-spline for $\Lop$. Then we claim that $\widehat{\widetilde{\varphi}}_j:=\abs{\widehat{\varphi}_j}^2$ defines the Fourier transform of a generalized B-spline for $\Lop^*\Lop$. An upper Riesz bound for $\widetilde{\varphi}_j$ can be found by using the norm inequality between $\ell_1$ and $\ell_2$: \begin{equation*} \sum_{\bm{k}\in\mathbb{Z}^d}\abs{\widehat\varphi_j(\bm{\omega}+2\pi (\mathbf{D}^T)^{-j}\bm{k})}^4 \leq \left(\sum_{\bm{k}\in\mathbb{Z}^d}\abs{\widehat\varphi_j(\bm{\omega}+2\pi (\mathbf{D}^T)^{-j}\bm{k})}^2 \right)^2 \leq B_j^2. \end{equation*} To verify the lower Riesz bound for $\widetilde{\varphi}_j$, we make use of the norm inequality \begin{equation}\label{eq:L*Lspline1} \sum_{\abs{\bm{k}}\leq M}\abs{\widehat\varphi_j(\bm{\omega}+2\pi (\mathbf{D}^T)^{-j}\bm{k})}^4 \geq C M^{-d}\left(\sum_{\abs{\bm{k}}\leq M}\abs{\widehat\varphi_j(\bm{\omega}+2\pi (\mathbf{D}^T)^{-j}\bm{k})}^2\right)^2, \end{equation} for an appropriately chosen $M>0$. Now, note that the decay condition of spline admissibility implies that for $\abs{\bm{\omega}}$ sufficiently large, there is a constant $C>0$ such that \begin{equation*} \abs{\widehat{L}(\bm{\omega})}^{-1} \leq C\abs{\bm{\omega}}^{-r} \end{equation*} This decay estimate on $\widehat{L}^{-1}$ combined with the lower Riesz bound for $\varphi_j$ gives \begin{align*} A_j &\leq \sum_{\bm{k}\in\mathbb{Z}^d}\abs{\widehat\varphi_j(\bm{\omega}+2\pi (\mathbf{D}^T)^{-j}\bm{k})}^2\\ &\leq \sum_{\abs{\bm{k}}\leq M}\abs{\widehat\varphi_j(\bm{\omega}+2\pi (\mathbf{D}^T)^{-j}\bm{k})}^2 + C\abs{\widehat{L}_{\rm{d},j}(\bm{\omega})}^2 M^{d-2r} \abs{\det(\mathbf{D})}^{2jr/d}, \end{align*} and hence \begin{equation}\label{eq:L*Lspline2} \sum_{\abs{\bm{k}}\leq M}\abs{\widehat\varphi_j(\bm{\omega}+2\pi (\mathbf{D}^T)^{-j}\bm{k})}^2 \geq A_j - C\abs{\widehat{L}_{\rm{d},j}(\bm{\omega})}^2 M^{d-2r} \abs{\det(\mathbf{D})}^{2jr/d}. \end{equation} Due to the fact that $2r>d$, we can always choose $M$ large enough to make the right hand side of \eqref{eq:L*Lspline2} positive. Using the estimate \eqref{eq:L*Lspline2} in \eqref{eq:L*Lspline1} establishes a lower Riesz bound for $\widetilde{\varphi}_j$ \end{proof} The Riesz basis property of the generalized B-splines for $\Lop^*\Lop$ imply that the $\Lop^*\Lop$-spline interpolants $\phi_j(\bm{x})$, given by \begin{equation}\label{eq:lagint} \widehat{\phi}_j(\bm{\omega}) = \abs{\det(\mathbf{D})}^j\frac{\abs{\widehat{\varphi}_j(\bm{\omega})}^2}{\sum_{\bm{k}\in \mathbb{Z}^d}\abs{\widehat{\varphi}_j(\bm{\omega}+2\pi (\mathbf{D}^T)^{-j} \bm{k})}^2}, \end{equation} are well-defined and also generate Riesz bases. Importantly, $\phi_j\in W_2^{\Lop}$ does not depend on the specific choice of the localization operator, as we can see from \begin{align*} \widehat{\phi}_j(\bm{\omega}) &= \abs{\det(\mathbf{D})}^j\frac{\abs{\widehat{L}_{\mathrm{d},j}(\bm{\omega})}^2\abs{\widehat{L}(\bm{\omega})}^{-2}}{\abs{\widehat{L}_{\mathrm{d},j}(\bm{\omega})}^2\sum_{\bm{k} \in \mathbb{Z}^d}\abs{\widehat{L}(\bm{\omega}+2\pi (\mathbf{D}^{T})^{-j}\bm{k})}^{-2}} \\ &= \abs{\det(\mathbf{D})}^j \frac{1}{1+\abs{\widehat{L}(\bm{\omega})}^{2}\sum_{\bm{k} \in \mathbb{Z}^d\backslash \{0\}}\abs{\widehat{L}(\bm{\omega}+2\pi (\mathbf{D}^{T})^{-j}\bm{k})}^{-2}}. \end{align*} These $\Lop^*\Lop$-spline interpolants play a key role in our wavelet construction, which we describe in the next section; however, for our approximation result, we are more interested in the related functions \begin{equation}\label{eq:lagmult} m_j(\bm{\omega}) = \frac{\abs{\widehat{L}(\bm{\omega})}^{-2}}{\sum_{\bm{k} \in \mathbb{Z}^d}\abs{\widehat{L}(\bm{\omega}+2\pi (\mathbf{D}^{T})^{-j}\bm{k})}^{-2}}, \end{equation} which are also needed for the decorrelation result, Theorem \ref{th:dc}. In order to bound the error of approximation from the spaces $V_j$, we apply the techniques developed in \cite{deboor94}. In that paper, the authors derive a characterization of certain potential spaces in terms of approximation by closed, shift-invariant subspaces of $L_2(\mathbb{R}^d)$. The same techniques can be applied in our situation, with a few modifications to account for smoothness being determined by different operator norms. The error in approximating a function $f\in L_2(\mathbb{R}^d)$ by a closed function space $X$ is denoted by \begin{equation*} E(f,X) := \min_{s\in X }\norm{f-s}_{L_2(\mathbb{R}^d)}, \end{equation*} and the approximation rate is given in terms of the density of the lattice in $\mathbb R^d$. The lattice determined by $\mathbf{D}^j$ has density proportional to $\abs{\det(\mathbf{D})}^{j/d}$, so we say that the multiresolution spaces $V_{j}$ provide approximation order $\tilde{r}$ if there is a constant $C>0$ such that \begin{equation*} E(f,V_{j}) \leq C\abs{\det(\mathbf{D})}^{j\tilde{r}/d} \norm{f}_{W_2^\Lop(\mathbb{R}^d)}, \end{equation*} for every $f\in W_2^\Lop(\mathbb{R}^d)$. \begin{theorem} For a spline-admissible pair $\Lop$ and $\mathbf{D}$ of order $r>d/2$, the multiresolution spaces $V_j$ provide approximation order $\tilde{r}\leq r$ if \begin{equation*} \abs{\det (\mathbf{D}^T)}^{-2j\tilde{r}/d} \frac{ 1-m_j(\bm{\omega}) }{1+\abs{\widehat{L}(\bm{\omega})}^2} \end{equation*} is bounded, independently of $j$, in $L_\infty((\mathbf{D}^T)^{-j}\Omega)$, where $\Omega=[-\pi,\pi]^d$. \end{theorem} \begin{proof} This result is a consequence of \cite[Theorem 4.3]{deboor94}. To show this let us introduce the notation $f_j(\cdot)=f(\mathbf{D}^{j}\cdot)$, which implies that $\widehat{f}_j=\abs{\det(\mathbf{D})}^{-j}\widehat{f} \circ (\mathbf{D}^T)^{-j}$, where $\circ$ denotes composition. The spaces $V_j$ are scaled copies of the integer shift-invariant spaces \begin{align*} V_{j}^j &:= \{ s(\mathbf{D}^j \cdot): s\in V_j \} \\ &= \left\{ \sum_{\bm{k}\in \mathbb{Z}^d} c[\bm{k}]\varphi_j (\mathbf{D}^j (\cdot- \bm{k})): c\in \ell_2(\mathbb Z^d)\right\}. \end{align*} We then write the error of approximating a function $f\in W_2^\Lop(\mathbb{R}^d)$ from $V_j$ in terms of approximation by $\widehat{V}_j^j$ as \begin{align*} E(f,V_j) &= \abs{\det (\mathbf{D})}^{j/2}E(f_j,V_j^j) \\ &= (2\pi)^{-d/2}\abs{\det (\mathbf{D})}^{j/2}E(\widehat{f}_j,\widehat{V}_j^j), \end{align*} where $\widehat{V}_j^j$ is composed of the Fourier transforms of functions in $V_j^j$. Separating this last term, we have \begin{equation}\label{eq:approx1} E(f,V_j) \leq (2\pi)^{-d/2}\abs{\det(\mathbf{D})}^{j/2}\left( E(\widehat{f}_j\chi_\Omega,\widehat{V}_j^j) +\norm{(1-\chi_\Omega)\widehat{f}_j}_2 \right), \end{equation} where $\chi_\Omega$ is the characteristic function of the set $\Omega$. We are now left with bounding both terms on the right-hand side of \eqref{eq:approx1}. First, we have \begin{align*} \norm{(1-\chi_\Omega)\widehat{f}_j}_2^2 &= \int_{\mathbb R^d\backslash \Omega} \abs{\widehat{f}_j(\bm{\omega})}^2 \rmd\bm{\omega} \\ &= \abs{\det(\mathbf{D})}^{-2j} \int_{\mathbb R^d\backslash \Omega} \abs{\widehat{f}((\mathbf{D}^T)^{-j}\bm{\omega})}^2 \frac{1+\abs{\widehat{L}((\mathbf{D}^T)^{-j}\bm{\omega})}^2}{1+\abs{\widehat{L}((\mathbf{D}^T)^{-j}\bm{\omega})}^2} \rmd\bm{\omega}, \end{align*} and since $\Lop$ is spline-admissible of order $r$ \begin{equation*} \norm{(1-\chi_\Omega)\widehat{f}_j}_2^2 \leq C_{\Lop} \abs{\det(\mathbf{D})}^{2jr/d-j} \norm{f}_{W_2^\Lop}^2. \end{equation*} Therefore \begin{equation}\label{eq:approx2} \abs{\det(\mathbf{D})}^{j/2}\norm{(1-\chi_\Omega)\widehat{f}_j}_2 \leq C_{\Lop}^{1/2} \abs{\det(\mathbf{D})}^{jr/d} \norm{f}_{W_2^\Lop}. \end{equation} In order to bound the remaining term, we need a formula for the projection of $\widehat{f}_j\chi_\Omega$ onto $V_j^j$. Notice that \begin{align*} 1-m_j((\mathbf{D}^T)^{-j}\cdot) &= 1-\frac{\abs{\widehat{\varphi}_j\circ (\mathbf{D}^T)^{-j}}^2}{\sum_{\bm{k}\in \mathbb Z^d}\abs{\widehat{\varphi}_j\circ (\mathbf{D}^T)^{-j}(\cdot - 2\pi \bm{k})}^2} \\ &= 1-\frac{\abs{\widehat{\varphi_j \circ \mathbf{D}^j}}^2}{\sum_{\bm{k}\in \mathbb Z^d}\abs{\widehat{\varphi_j \circ \mathbf{D}^j}(\cdot - 2\pi \bm{k})}^2} , \end{align*} so we apply \cite[Theorem 2.20]{deboor94} to get \begin{align*} E(\widehat{f}_j\chi_\Omega,\widehat{V}_j^j)^2 &= \int_{\Omega} \abs{\widehat{f}_j}^2 (1-m_j((\mathbf{D}^T)^{-j}\cdot) ) \\ &= \abs{\det(\mathbf{D})}^{-2j} \int_{\Omega} \abs{\widehat{f} \circ (\mathbf{D}^T)^{-j}}^2 (1-m_j((\mathbf{D}^T)^{-j}\cdot) ). \end{align*} Now, changing variables gives \begin{align*} E(\widehat{f}_j\chi_\Omega,\widehat{V}_j^j)^2 &= \abs{\det(\mathbf{D})}^{-j} \int_{(\mathbf{D}^T)^{-j} \Omega} \abs{\widehat{f} }^2 \left(1+\abs{\widehat{L}}^2\right) \frac{1-m_j}{1+\abs{\widehat{L}}^2} \\ &\leq \abs{\det(\mathbf{D})}^{-j} \norm{f}_{W_2^\Lop}^2 \norm{ \frac{1-m_j}{1+\abs{\widehat{L}}^2}}_{L_\infty((\mathbf{D}^T)^{-j}\Omega)}. \end{align*} Applying our assumption on $(1-m_j)$, we have \begin{equation}\label{eq:approx3} \abs{\det(\mathbf{D})}^{j/2} E(\widehat{f}_j\chi_\Omega,\widehat{V}_j^j) \leq C \abs{\det(\mathbf{D})}^{j \tilde{r}/d} \norm{f}_{W_2^\Lop}. \end{equation} Substituting the estimates \eqref{eq:approx2} and \eqref{eq:approx3} into \eqref{eq:approx1} yields the result. \end{proof} Concerning this theorem, an important point is that it describes the approximation properties of the MRA entirely in terms of the operator; i.e., the guaranteed approximation rates are independent of how one chooses the generalized B-splines $\varphi_j$ for the multiresolution spaces $V_j$. \section{Operator-Like Wavelets and Riesz Bases}\label{sec:wave} Using the non-stationary MRA defined in the previous section, we define the scale of wavelet spaces $W_j$ by the relationship \begin{equation*} V_{j} = V_{j+1} \oplus W_{j+1}; \end{equation*} i.e., $W_{j+1}$ is the orthogonal complement of $V_{j+1}$ in $V_j$. Our goal in this section is to define Riesz bases for these spaces and for $L_2(\mathbb{R}^d)$. To begin, let us define the functions \begin{equation*} \psi_{j+1} := \Lop^*\phi_{j}, \end{equation*} which we claim generate Riesz bases for the wavelet spaces, under mild conditions on the operator $\Lop$. First, note that $\psi_{j+1}$ is indeed in $V_{j}$, because its Fourier transform $\widehat{\psi}_{j+1}$ is a periodic multiple of $\widehat{\varphi}_j$, and thus \begin{equation}\label{eq:oplikewavfourier} \widehat{\psi}_{j+1}(\bm{\omega}) = \abs{\det(\mathbf{D})}^j \frac{\widehat{L}_{\mathrm{d},j}(\bm{\omega})^*}{ \sum_{\bm{k}\in \mathbb{Z}^d}\abs{\widehat{\varphi}_j(\bm{\omega}+2\pi (\mathbf{D}^T)^{-j} \bm{k})}^2 } \widehat{\varphi}_j(\bm{\omega}). \end{equation} A direct implication of our wavelet construction is the following property. \begin{property} The wavelet function $\psi_{j+1}$ behaves like a multiscale version of the underlying operator $\Lop$ in the sense that, for any $f\in W_2^{\Lop}$, we have $f*\psi_{j+1}=\Lop^*(f*\phi_j)$. Hence, in the case where $\phi_j$ is a lowpass filter, $\{\Lop^*(f*\phi_j)\}_{j\in \mathbb{Z}}$ corresponds to a multiscale representation of $\Lop^*f$. \end{property} The next few results focus on showing that the $\mathbf{D}^j\mathbb{Z}^d \setminus \mathbf{D}^{j+1}\mathbb{Z}^d$ shifts of $\psi_{j+1}$ are orthogonal to $V_{j+1}$ and generate a Riesz basis of $W_{j+1}$. \begin{proposition}\label{prop:orthog_v_w} The wavelets $\{\psi_{j+1}(\cdot-\mathbf{D}^j\bm{k})\}_{\bm{k}\in{\mathbb{Z}^d\backslash \mathbf{D}\mathbb{Z}^d}}$ are orthogonal to the space $V_{j+1}$. \end{proposition} \begin{proof} It suffices to show $\langle \varphi_{j+1}, \psi_{j+1}(\cdot-\mathbf{D}^j\bm{k}) \rangle=0$ for every $\bm{k} \in \mathbb{Z}^d\backslash \mathbf{D}\mathbb{Z}^d$. From \eqref{eq:oplikewavfourier}, we have \begin{align*} \langle \varphi_{j+1}, \psi_{j+1}(\cdot-\mathbf{D}^j\bm{k}) \rangle &= \int_{\mathbb{R}^d} \widehat{\varphi}_{j+1}(\bm{\omega})e^{i\bm{\omega} \cdot\mathbf{D}^j\bm{k}} \widehat{L}(\bm{\omega})\widehat{\phi}_j(\bm{\omega}) \rmd\bm{\omega} \\ &= \int_{\mathbb{R}^d} \widehat{L}_{\mathrm{d},j+1}(\bm{\omega})e^{i\bm{\omega} \cdot\mathbf{D}^j\bm{k}} \widehat{\phi}_j(\bm{\omega}) \rmd\bm{\omega}. \end{align*} Now let $\Omega$ be a fundamental domain for the lattice $2\pi (\mathbf{D}^T)^{-j}\mathbb{Z}^d$. Then \begin{align*} \langle \varphi_{j+1}, \psi_{j+1}(\cdot-\mathbf{D}^j\bm{k}) \rangle &= \int_{\Omega} \widehat{L}_{\mathrm{d},j+1}(\bm{\omega})e^{i\bm{\omega} \cdot\mathbf{D}^j\bm{k}} \sum_{\bm{n} \in \mathbb{Z}^d}\widehat{\phi}_j(\bm{\omega}-2\pi (\mathbf{D}^T)^{-j} \bm{n}) \rmd\bm{\omega} \\ &= \abs{\det(\mathbf{D})}^j \int_{\Omega} \widehat{L}_{\mathrm{d},j+1}(\bm{\omega})e^{i\bm{\omega} \cdot\mathbf{D}^j\bm{k}} \rmd\bm{\omega}. \end{align*} From Definition \ref{def:spadm}, we know that $\widehat{L}_{\mathrm{d},j+1}$ has a series representation of the form $\sum_{\bm{n} \in \mathbb{Z}^d} p_{j+1}[\bm{n}]e^{i\bm{\omega}\cdot \mathbf{D}^{j+1} \bm{n}}$, so \begin{align*} \langle \varphi_{j+1}, \psi_{j+1}(\cdot-\mathbf{D}^j\bm{k}) \rangle &= \abs{\det(\mathbf{D})}^j \sum_{\bm{n} \in \mathbb{Z}^d} p_{j+1}[\bm{n}] \int_{\Omega} e^{i\bm{\omega}\cdot \mathbf{D}^j \left(\mathbf{D}\bm{n} + \bm{k} \right)} \rmd\bm{\omega}\\ &= \sum_{\bm{n} \in \mathbb{Z}^d} p_{j+1}[\bm{n}] \int_{[0,2\pi]^d} e^{i\bm{\omega}\cdot \left(\mathbf{D}\bm{n} + \bm{k} \right)} \rmd\bm{\omega}. \end{align*} Since $\bm{k} \notin \mathbf{D} \mathbb{Z}^d$, we see that $\mathbf{D}\bm{n} + \bm{k} \neq 0$ for any $\bm{n}$, and this implies that $\langle \varphi_{j+1}, \psi_{j+1}(\cdot-\mathbf{D}^j\bm{k}) \rangle=0$. \end{proof} In order to prove that the $\mathbf{D}^j\mathbb{Z}^d \setminus \mathbf{D}^{j+1}\mathbb{Z}^d$ shifts of $\psi_{j+1}$ form a Riesz basis of the wavelet space $W_{j+1}$, we introduce notation that will help us formulate the problem as a shift-invariant one. In the following definition, we use the fact that there is a set of vectors \begin{equation*} \left\{\bm{e}_{l}\in \mathbb{Z}^d: l=0,..,\abs{\det(\mathbf{D})}-1\right\} \end{equation*} such that \begin{equation*} \bigcup_{l=0}^{\abs{\det(\mathbf{D})}-1} \left(\mathbf{D}^j\bm{e}_{l} + \mathbf{D}^{j+1}\mathbb{Z}^d\right) = \mathbf{D}^{j} \mathbb{Z}^d. \end{equation*} \begin{definition} For every $j\in\mathbb{Z}$ and every $l\in \{1,\dots,\abs{\det(\mathbf{D})}-1\}$, we define the wavelets \begin{equation*} \psi_{j+1}^{(l)} (\bm{x}) := \psi_{j+1} (\bm{x}-\mathbf{D}^j\bm{e}_{l}), \end{equation*} and we define the collections \begin{equation*} \Psi := \Psi_{j+1} := \left\{\psi^{(l)}_{j+1}\right\}_{l=1}^{\abs{\det(\mathbf{D})}-1}. \end{equation*} \end{definition} In the following, necessary and sufficient conditions on the operator $\Lop$ are given which guarantee that $\Psi_{j+1}$ generates a Riesz basis of $W_{j+1}$. The technique used is called fiberization, and it can be applied to characterize finitely generated shift-invariant spaces \cite{ron95}. In this setting, a collection of functions defines a Gramian matrix, and the property of being a Riesz basis is equivalent to the Gramian having bounded eigenvalues. In our situation, the Gramian for $\Psi$ is \begin{align*} G_{\Psi}(\bm{\omega}) &= \abs{\det(\mathbf{D})}^{-j-1} \left( \sum_{\bm{\beta} \in 2\pi (\mathbf{D}^T)^{-j-1}\mathbb{Z}^d} \widehat{\psi^{(k)}_{j+1}}\left(\bm{\omega}+\bm{\beta} \right) \widehat{\psi^{(l)}_{j+1}}\left(\bm{\omega}+\bm{\beta} \right)^* \right)_{k,l}\\ &= \abs{\det(\mathbf{D})}^{-j-1} \left( \sum_{\bm{\beta} \in 2\pi (\mathbf{D}^T)^{-j-1}\mathbb{Z}^d} e^{- i \mathbf{D}^{j}(\bm{e}_k-\bm{e}_l)\cdot (\bm{\omega}+\bm{\beta})} \abs{\widehat{\psi}_{j+1}(\bm{\omega}+\bm{\beta})}^2 \right)_{k,l}, \end{align*} where $k$ and $l$ range from $1$ to $\abs{\det(\mathbf{D})}-1$. The normalization factor $\abs{\det(\mathbf{D})}^{-j-1}$ accounts for scaling of the lattice. Let us denote the largest and smallest eigenvalues of $G_{\Psi}(\bm{\omega})$ by $\Lambda(\bm{\omega})$ and $\lambda(\bm{\omega})$, respectively. Then the collection $\Psi$ generates a Riesz basis if and only if $\Lambda$ and $1/\lambda$ are essentially bounded (cf. \cite{ron95} Theorem 2.3.6). To simplify this matrix without changing the eigenvalues, we apply the similarity transformation $T(\bm{\omega})^{-1}G_{\Psi}(\bm{\omega})T(\bm{\omega})$, where $T$ is the square diagonal matrix with diagonal entry $e^{-i\mathbf{D}^{j}\bm{e}_l\cdot\bm{\omega}}$ in row $l$. This transformation multiplies column $l$ of $G_\Psi$ by $e^{-i\mathbf{D}^{j}\bm{e}_l\cdot\bm{\omega}}$ and row $k$ of $G_\Psi$ by $e^{i\mathbf{D}^{j}\bm{e}_k\cdot\bm{\omega}}$. Since the eigenvalues are unchanged, let us call this new matrix $G_\Psi$ as well. We then have \begin{equation*} G_\Psi (\bm{\omega}) = \abs{\det(\mathbf{D})}^{-j-1}\left( \sum_{\bm{\beta}\in 2\pi (\mathbf{D}^T)^{-j-1}\mathbb{Z}^d} e^{- i \mathbf{D}^{j}(\bm{e}_k-\bm{e}_l)\cdot \bm{\beta}} \abs{\widehat{\psi}_{j+1}(\bm{\omega}+\bm{\beta})}^2 \right)_{k,l}. \end{equation*} Using the fact that $\bigcup_m \left(\bm{e}_m+\mathbf{D}^T\mathbb{Z}^d\right)=\mathbb{Z}^d$ and the notation \begin{equation}\label{eq:rieszbds_cmw} c(m;\bm{\omega}) := \abs{\det(\mathbf{D})}^{-j-1}\sum_{\bm{\beta}\in 2\pi (\mathbf{D}^T)^{-j} \mathbb{Z}^d} \abs{\widehat{\psi}_{j+1}(\bm{\omega}+2\pi (\mathbf{D}^T)^{-j-1} \bm{e}_m+\bm{\beta})}^2, \end{equation} we write \begin{equation*} G_\Psi(\bm{\omega}) = \left( \sum_{m=0}^{\abs{\det(\mathbf{D})}-1} c(m;\bm{\omega}) e^{-2\pi i (\bm{e}_k-\bm{e}_l)\cdot (\mathbf{D}^T)^{-1}\bm{e}_m} \right)_{k,l}. \end{equation*} \begin{definition} Let $\mathbf{H}$ be the $\abs{\det(\mathbf{D})} \times \abs{\det(\mathbf{D})}$ matrix \begin{equation*} \mathbf{H} := \abs{\det(\mathbf{D})}^{-1/2} \left( e^{2\pi i\bm{e}_m\cdot (\mathbf{D}^T)^{-1}\bm{e}_k} \right)_{k,m}, \end{equation*} which is the complex conjugate of the discrete Fourier transform matrix for the lattice generated by $\mathbf{D}^T$ \cite{vaidyanathan90}. Here, $k$ and $m$ range over the index set $\mathcal{M}:=\{0,\dots,\abs{\det(\mathbf{D})}-1\}$. Also, define $\mathbf{H}_0$ to be the submatrix obtained by removing column $0$ from $\mathbf{H}$. \end{definition} \begin{lemma}\label{lem:eig_values} The minimum and maximum eigenvalues $\lambda(\bm{\omega}),\Lambda(\bm{\omega})$ of the Gramian matrix $G_{\Psi}(\bm{\omega})$ satisfy the following properties: \begin{enumerate}[(i)] \item ${\displaystyle \lambda(\bm{\omega}) \geq \abs{\det(\mathbf{D})} \min_{m\in\mathcal{M}} c(m;\bm{\omega}) \quad \text{and} \quad \Lambda(\bm{\omega}) \leq \abs{\det(\mathbf{D})} \max_{m\in\mathcal{M}} c(m;\bm{\omega}) }$ \item There is a constant $C>0$ such that \begin{equation*} \lambda(\bm{\omega}) \geq C \abs{\det(\mathbf{D})} \max_{m_0(\bm{\omega}) \in\mathcal{M} }\min_{ m\in \mathcal{M} \setminus \{m_0(\bm{\omega})\} } c(m;\bm{\omega}). \end{equation*} \item If for any fixed $\bm{\omega}\in\mathbb{R}^d$, there exist distinct $m_1(\bm{\omega}),m_2(\bm{\omega})\in \mathcal{M}$ such that $c(m_1(\bm{\omega});\bm{\omega})=c(m_2(\bm{\omega});\bm{\omega})=0$, then $\lambda(\bm{\omega})=0$. \end{enumerate} \end{lemma} \begin{proof} The Gramian matrix $G_{\Psi}$ can be written as \begin{equation*} G_\Psi(\bm{\omega}) = \abs{\det(\mathbf{D})} \mathbf{H}_0^* \mathcal{D}(\bm{\omega}) \mathbf{H}_0, \end{equation*} where $\mathcal{D}(\bm{\omega})$ is the $\abs{\det(\mathbf{D})} \times \abs{\det(\mathbf{D})}$ diagonal matrix with entry $c(m;\bm{\omega})$ in column $m$: \begin{align*} \abs{\det(\mathbf{D})} \mathbf{H}_0^* \mathcal{D}(\bm{\omega}) \mathbf{H}_0 &= \left( e^{-2\pi i\bm{e}_k\cdot (\mathbf{D}^T)^{-1}\bm{e}_n} \right)_{k,n} \left(c(m;\bm{\omega})\right)_{n,m} \left( e^{2\pi i\bm{e}_l\cdot (\mathbf{D}^T)^{-1}\bm{e}_m} \right)_{m,l} \\ &= \left( c(m;\bm{\omega})e^{-2\pi i\bm{e}_k\cdot (\mathbf{D}^T)^{-1}\bm{e}_m} \right)_{k,m} \left( e^{2\pi i\bm{e}_l\cdot (\mathbf{D}^T)^{-1}\bm{e}_m} \right)_{m,l} \\ &= \left(\sum_{m=0}^{\abs{\det(\mathbf{D})}-1} c(m;\bm{\omega})e^{-2\pi i (\bm{e}_k -\bm{e}_l )\cdot (\mathbf{D}^T)^{-1}\bm{e}_m} \right)_{k,l}. \end{align*} Since $\mathcal{D}(\bm{\omega})$ has non-negative entries, we write this as \begin{equation*} G_\Psi(\bm{\omega}) = \abs{\det(\mathbf{D})} (\mathcal{D}(\bm{\omega})^{1/2} \mathbf{H}_0)^* (\mathcal{D}(\bm{\omega})^{1/2} \mathbf{H}_0). \end{equation*} Now consider the quadratic form \begin{align*} \bm{\alpha}^*G_\Psi(\bm{\omega})\bm{\alpha} &= \abs{\det(\mathbf{D})} (\mathcal{D}(\bm{\omega})^{1/2} \mathbf{H}_0\bm{\alpha})^* (\mathcal{D}(\bm{\omega})^{1/2} \mathbf{H}_0\bm{\alpha})\\ &= \abs{\det(\mathbf{D})} \abs{\mathcal{D}(\bm{\omega})^{1/2} \mathbf{H}_0\bm{\alpha}}^2, \end{align*} where $\bm{\alpha}\in\mathbb{C}^{\abs{\det(\mathbf{D})}-1}$. Since $\mathbf{H}_0$ is an isometry, $\abs{\mathbf{H}_0 \bm{\alpha}}=\abs{\bm{\alpha}}$, and we immediately verify {\it (i)}. To prove {\it (ii)}, we first identify the range of $\mathbf{H}_0$. By the Fredholm alternative, a vector is in the range of $\mathbf{H}_0$ if and only if it is orthogonal to the null space of $\mathbf{H}_0^*$. Since $\mathbf{H}^*$ is a unitary matrix and its first row is a constant multiple of $(1,1,\dots,1)^T$, the range of $\mathbf{H}_0$ consists of vectors that are orthogonal to $(1,1,\dots,1)^T$. Therefore \begin{align*} \lambda(\bm{\omega}) & = \abs{\det(\mathbf{D})} \min_{\substack{\bm{\alpha}\in\mathbb{C}^{\abs{\det(\mathbf{D})}-1} \\ \abs{\bm{\alpha}}=1}} \abs{\mathcal{D}(\bm{\omega})^{1/2} \mathbf{H}_0\bm{\alpha}}^2\\ & = \abs{\det(\mathbf{D})} \min_{\substack{\bm{\alpha}\in\mathbb{C}^{\abs{\det(\mathbf{D})}} \\ \abs{\bm{\alpha}}=1 \\ \bm{\alpha}\perp (1,1,\dots,1)^T}} \abs{\mathcal{D}(\bm{\omega})^{1/2} \bm{\alpha}}^2 \\ & = \abs{\det(\mathbf{D})} \min_{\substack{\bm{\alpha}\in\mathbb{C}^{\abs{\det(\mathbf{D})}} \\ \abs{\bm{\alpha}}=1 \\ \bm{\alpha}\perp (1,1,\dots,1)^T}} \sum_{m\in \mathcal{M}} \abs{\alpha_m}^2 c(m;\bm{\omega}), \end{align*} where in the last equation, we use the notation $\bm{\alpha} = (\alpha_0,\dots,\alpha_{\abs{\det(\mathbf{D})}-1})$. Then a lower bound is given by \begin{align*} \lambda(\bm{\omega}) &\geq \abs{\det(\mathbf{D})} \max_{m_0(\bm{\omega}) \in\mathcal{M} } \min_{\substack{\bm{\alpha}\in\mathbb{C}^{\abs{\det(\mathbf{D})}} \\ \abs{\bm{\alpha}}=1 \\ \bm{\alpha}\perp (1,1,\dots,1)^T}} \sum_{m\in \mathcal{M}\setminus \{m_0(\bm{\omega})\} } \abs{\alpha_m}^2 c(m;\bm{\omega}) \\ &\geq \abs{\det(\mathbf{D})} \max_{m_0(\bm{\omega}) \in\mathcal{M} } \min_{\substack{\bm{\alpha}\in\mathbb{C}^{\abs{\det(\mathbf{D})}} \\ \abs{\bm{\alpha}}=1 \\ \bm{\alpha}\perp (1,1,\dots,1)^T}} \left( \min_{m\in \mathcal{M} \setminus \{m_0(\bm{\omega})\}} c(m;\bm{\omega})\right) \sum_{m\in \mathcal{M}\setminus \{m_0(\bm{\omega})\} } \abs{\alpha_m}^2 \\ & = \abs{\det(\mathbf{D})} \max_{m_0(\bm{\omega}) \in\mathcal{M} } \left(\min_{m\in \mathcal{M} \setminus \{m_0(\bm{\omega})\}} c(m;\bm{\omega})\right) \left( \min_{\substack{\bm{\alpha}\in\mathbb{C}^{\abs{\det(\mathbf{D})}} \\ \abs{\bm{\alpha}}=1 \\ \bm{\alpha}\perp (1,1,\dots,1)^T}} \sum_{m\in \mathcal{M}\setminus \{m_0(\bm{\omega})\} } \abs{\alpha_m}^2 \right). \end{align*} Notice that none of the standard unit vectors \begin{equation*} \{(1,0,\dots,0),(0,1,0,\dots,0),\dots,(0,\dots,0,1)\} \end{equation*} are orthogonal to $(1,1,\dots,1)^T$, so there is a constant $C>0$ such that \begin{equation*} \min_{m_0(\bm{\omega})\in \mathcal{M}}\min_{\substack{\bm{\alpha}\in\mathbb{C}^{\abs{\det(\mathbf{D})}} \\ \abs{\bm{\alpha}}=1 \\ \bm{\alpha}\perp (1,1,\dots,1)^T}} \sum_{m\in\mathcal{M}\setminus \{m_0(\bm{\omega})\}}\abs{\alpha_m}^2 = C. \end{equation*} We now use this constant to provide a lower bound for $\lambda(\bm{\omega})$: \begin{equation*} \lambda(\bm{\omega}) \geq C \abs{\det(\mathbf{D})} \max_{m_0(\bm{\omega}) \in\mathcal{M} }\min_{m\in \mathcal{M} \setminus \{m_0(\bm{\omega})\}} c(m;\bm{\omega}). \end{equation*} Finally, for {\it (iii)}, suppose that there are distinct $m_1(\bm{\omega}),m_2(\bm{\omega})\in \mathcal{M}$ such that $c(m_1(\bm{\omega});\bm{\omega})=c(m_2(\bm{\omega});\bm{\omega})=0$. Then define the vector $\bm{\alpha}=(\alpha_1,\dots,\alpha_{\abs{\det(\mathbf{D})}-1})\in \mathbb{C}^{\abs{\det(\mathbf{D})}}$ such that $\alpha_{m_1}=1/\sqrt{2}$, $\alpha_{m_2}=-1/\sqrt{2}$, and all other entries are zero. This vector is in the range of $\mathbf{H}_0$, and $\mathcal{D}(\bm{\omega})^{1/2} \bm{\alpha}=0$. Therefore $\lambda(\bm{\omega})=0$. \end{proof} \begin{lemma}\label{lem:rb} The collection $\Psi$ generates a Riesz basis if and only if no two of the functions $c(m;\bm{\omega})$ are zero for the same $\bm{\omega}$. \end{lemma} \begin{proof} Let $\Omega_j$ be a fundamental domain for the lattice $2\pi (\mathbf{D}^T)^{-j}\mathbb{Z}^d$, and let $\overline{\Omega}_j$ denote its closure. For the reverse direction, we must show that there is a uniform lower bound of $\lambda(\bm{\omega})$ over $\overline{\Omega}_j$. By Lemma \ref{lem:eig_values}, it suffices to provide a lower bound for \begin{equation}\label{eq:riesz_uniform_fd} \max_{m_0(\bm{\omega}) \in \mathcal{M}} \min_{m\in \mathcal{M}\setminus \{m_0(\bm{\omega})\}} c(m;\bm{\omega}). \end{equation} Based on \eqref{eq:oplikewavfourier} and \eqref{eq:rieszbds_cmw}, we verify that \begin{equation}\label{eq:coeff_fcn_rb} c(m;\bm{\omega}) = \abs{\det(\mathbf{D})}^{j-1} \frac{\abs{ \widehat{L}_{\mathrm{d},j}(\bm{\omega}+2\pi (\mathbf{D}^T)^{-j-1} \bm{e}_m)}^2}{\sum_{\bm{k}\in \mathbb{Z}^d} \abs{\widehat{\varphi}_j(\bm{\omega}+2\pi (\mathbf{D}^T)^{-j-1} \bm{e}_m+2\pi (\mathbf{D}^T)^{-j} \bm{k})}^2}. \end{equation} Note that the numerator is a continuous function, and the denominator is bounded away from zero, due to the Riesz basis condition on $\varphi_j$. Hence, $c(m;\bm{\omega})=0$ at a point $\bm{\omega}$ if and only if $\widehat{L}_{\mathrm{d},j}(\bm{\omega}+2\pi (\mathbf{D}^T)^{-j-1} \bm{e}_m)=0$. Let us define the continuous function \begin{equation}\label{eq:cont_lower_bnd} F(\bm{\omega}) := \max_{m_0(\bm{\omega}) \in \mathcal{M}} \min_{m\in \mathcal{M}\setminus \{m_0(\bm{\omega})\}} \abs{ \widehat{L}_{\mathrm{d},j}(\bm{\omega}+2\pi (\mathbf{D}^T)^{-j-1} \bm{e}_m)}^2. \end{equation} Since no two functions $c(m;\bm{\omega})$ are zero at any point $\bm{\omega}$, $F$ is positive on $\overline{\Omega}_j$. Due to the compactness of this set, there is a constant $C>0$ such that $F(\bm{\omega})>C$ on $\overline{\Omega}_j$. Since $F$ is bounded away from zero, \eqref{eq:riesz_uniform_fd} is as well. The forward direction follows from {\it (iii)} of Lemma \ref{lem:eig_values}. \end{proof} \begin{lemma}\label{lem:rb_w} If $\Psi_{j+1}$ generates a Riesz basis, then it provides a Riesz basis for $W_{j+1}$. \end{lemma} \begin{proof} We verify this fact by comparing the bases \begin{align*} \Psi_{j+1}' & := \{\varphi_{j+1}\} \bigcup \Psi_{j+1}, \\ \Phi_{j} & := \{\varphi_{j}(\cdot-\mathbf{D}^{j}\bm{e}_m)\}_{m\in\mathcal{M}} \end{align*} for shift invariant spaces on the lattice $\mathbf{D}^{j+1}\mathbb{Z}^d$. The $\mathbf{D}^{j}\mathbb{Z}^d$ shifts of $\varphi_j$ are a Riesz basis for $V_j$, or, equivalently, the $\mathbf{D}^{j+1}\mathbb{Z}^d$ shifts of the elements of $\Phi_{j}$ are a Riesz basis of $V_j$. This basis has $\abs{\det(\mathbf{D})}$ elements, and any other basis must have the same number of elements. The collections $\Psi_{j+1}$ and $\{\varphi_{j+1}\}$ generate Riesz bases, and both are contained in $V_j$. These bases are orthogonal, as was shown in Proposition \ref{prop:orthog_v_w}. Therefore $\Psi_{j+1}'$ generates a Riesz basis for a subspace of $V_{j}$, and $\Psi_{j+1}$ generates a Riesz basis for a subspace of $W_{j+1}$. The fact that $\Psi_{j+1}'$ has $\abs{\det(\mathbf{D})}$ elements implies that $\Psi_{j+1}'$ provides a Riesz basis for $V_j$ (cf. \cite[Theorem 2.26]{deboor93} and \cite{aldroubi95}), and the result follows. \end{proof} In Lemmas \ref{lem:eig_values} and \ref{lem:rb}, we saw how $\Psi$ generating a Riesz basis depends on the zeros of the functions $c(m;\cdot)$. From \eqref{eq:coeff_fcn_rb}, it is clear that the zeros of $c(m;\cdot)$ coincide with the zeros of a shifted version of $\widehat{L}_{\mathrm{d},j}$. In order to interpret the Riesz basis conditions in terms of the operator $\Lop$, we note that the zeros of $\widehat{L}_{\mathrm{d},j}$ are precisely the periodized zeros of $\widehat{L}$. Let us denote the zero set of the symbol $\widehat{L}$ as \begin{equation*} \mathcal N := \{\bm{p}\in{\mathbb R^d} : \widehat{L}(\bm{p})=0\}, \end{equation*} and for each scale $j$ and each $m=0,\dots, \abs{\det(\mathbf{D})}-1$, let us define the periodized sets \begin{equation*} \mathcal{N}_j^{(m)} := \left\{ \bm{p}-2\pi (\mathbf{D}^T)^{-j-1} \bm{e}_m+2\pi (\mathbf{D}^T)^{-j}\bm{k} : \bm{p}\in \mathcal{N}, \bm{k}\in \mathbb{Z}^d \right\}. \end{equation*} Note that $\mathcal{N}_j^{(m)}$ is the zero set of $\widehat{L}_{\mathrm{d},j}(\cdot + 2\pi (\mathbf{D}^T)^{-j-1} \bm{e}_m )$, and hence it is also the zero set of $c(m,\bm{\omega})$. \begin{theorem}\label{th:rb} Let $j\in \mathbb{Z}$ be an arbitrary scale. Then the family of functions \begin{equation*} \Psi_{j+1} = \left\{\psi^{(m)}_{j+1}\right\}_{m=1}^{\abs{\det(\mathbf{D})}-1} \end{equation*} generates a Riesz basis of $W_{j+1}$ if and only if the sets $\mathcal N_j^{(m)}$ satisfy \begin{equation}\label{eq:suffcond} \mathcal N_j^{(0)}\cap\mathcal N_j^{(m)} = \emptyset \end{equation} for each $1\leq m\leq \abs{\det(\mathbf{D})}-1$. \end{theorem} Our wavelet construction is intended to be general so that we may account for a large collection of operators. As a consequence of this generality, we cannot conclude that our wavelet construction always produces a Riesz basis of $L_2(\mathbb{R}^d)$. Here, we impose additional conditions on $\Lop$ and multiply $\psi_{j+1}$ by an appropriate normalization factor to ensure that a Riesz basis is produced. In order to preserve generality, we focus on the fine scale wavelet spaces and include a multiresolution space $V_{j_0+1}$ in our Riesz basis. \begin{theorem}\label{th:rb2} Let $\Lop$ be a spline admissible operator of order $r$, and suppose that there exist $\omega_0>0$ and constants $C_1,C_2>0$ such that the symbol $\widehat{L}$ satisfies \begin{equation*} C_1 \abs{\bm{\omega}}^{r} \leq \abs{\widehat{L}(\bm{\omega})} \leq C_2 \abs{\bm{\omega}}^r \end{equation*} for $\abs{\bm{\omega}} \geq \omega_0$. Then there is an integer $j_0$ such that the collection \begin{equation}\label{eq:rb} \left\{\varphi_{j_0+1}(\cdot-\beta)\right\}_{\beta \in \mathbf{D}^{j_0+1}\mathbb{Z}^d} \bigcup_{j\leq j_0} \left\{ \abs{\det(\mathbf{D})}^{(r/d-1/2)j}\psi_{j+1}(\cdot-\beta)\right\}_{\beta \in \mathbf{D}^{j}\mathbb{Z}^d\backslash\mathbf{D}^{j+1}\mathbb{Z}^d } \end{equation} forms a Riesz basis of $L_2(\mathbb{R}^d)$. \end{theorem} \begin{proof} Let $j_0$ be an integer for which $\omega_0<\pi/4 \abs{\det(\mathbf{D})}^{-j_0/d}$. Considering Lemma \ref{lem:eig_values}, the Riesz bounds for the wavelet spaces depend on the functions $c(m;\bm{\omega})$ of \eqref{eq:rieszbds_cmw}. A Fourier domain formula for the wavelet $\psi_{j+1}$ is \begin{equation*} \widehat{\psi}_{j+1} = \abs{ \text{det} (\mathbf{D}) }^j \frac{\widehat{L}(\bm{\omega})^{-1}}{\sum_{\bm{k}\in\mathbb{Z}^d} \abs{\widehat{L}(\bm{\omega}+2\pi (\mathbf{D}^T)^{-j}\bm{k})}^{-2}}, \end{equation*} so we have \begin{align*} c(m;\bm{\omega}) &= \abs{\det(\mathbf{D})}^{-j-1}\sum_{\bm{\beta}\in 2\pi (\mathbf{D}^T)^{-j} \mathbb{Z}^d} \abs{\widehat{\psi}_{j+1}(\bm{\omega}+2\pi (\mathbf{D}^T)^{-j-1} \bm{e}_m+\bm{\beta})}^2\\ &= \abs{ \text{det} (\mathbf{D}) }^{j-1} \sum_{\bm{\beta}\in 2\pi (\mathbf{D}^T)^{-j} \mathbb{Z}^d} \abs{ \frac{\widehat{L}(\bm{\omega}+2\pi (\mathbf{D}^T)^{-j-1}\bm{e}_m +\bm{\beta})^{-1}}{\sum_{\bm{k}\in\mathbb{Z}^d} \abs{\widehat{L}(\bm{\omega}+2\pi (\mathbf{D}^T)^{-j-1}\bm{e}_m +2\pi (\mathbf{D}^T)^{-j}\bm{k})}^{-2}} }^2. \end{align*} We now need upper and lower bounds on the terms \begin{equation}\label{eq:psi_sum} \sum_{\bm{\beta}\in 2\pi (\mathbf{D}^T)^{-j} \mathbb{Z}^d} \abs{\frac{\widehat{L}(\bm{\omega}+2\pi (\mathbf{D}^T)^{-j-1} \bm{e}_m+\bm{\beta})^{-1}}{\sum_{\bm{k} \in \mathbb{Z}^d}\abs{\widehat{L}(\bm{\omega}+2\pi (\mathbf{D}^T)^{-j-1} \bm{e}_m+2\pi (\mathbf{D}^T)^{-j}\bm{k})}^{-2}}}^2. \end{equation} Recall that the upper bound should be uniform across all values of $m$; however, for the lower bound, it is sufficient to consider only $\abs{\det(\mathbf{D})}-1$ of the functions $c(m;\bm{\omega})$. Define the lattices \begin{equation*} X_j(m,\bm{\omega}) := \left\{ \bm{x}_{\bm{k}}=\bm{\omega}+2\pi (\mathbf{D}^T)^{-j-1} \bm{e}_m+2\pi (\mathbf{D}^T)^{-j}\bm{k}: \bm{k}\in\mathbb{Z}^d \right\}. \end{equation*} The value of \eqref{eq:psi_sum} depends on position of $X_j(m,\bm{\omega})$ with respect to the origin, as well as two density parameters. Let us introduce the notation $h_j$ for the fill distance and $q_j$ for the separation radius of $X_j(m,\bm{\omega})$. Since each lattice $X_j(m,\bm{\omega})$ is a translation of $X_j(0,\bm{0})$, the quantities $h_j$ and $q_j$ are independent of $m$ and $\bm{\omega}$, and they are defined as \begin{align*} h_j &:= \sup_{\bm{y} \in \mathbb{R}^d} \inf_{\bm{x}\in X_j(0,\bm{0})} \abs{\bm{y}-\bm{x}}\\ q_j &:= \frac{1}{2}\inf_{ \substack{\bm{x}, \bm{x}' \in X_j(0,\bm{0}) \\ \bm{x}\neq \bm{x}' }} \abs{\bm{x}-\bm{x}'}. \end{align*} Given the structure of the matrix $\mathbf{D}$, we can compute \begin{align*} h_j &= 2\pi \abs{\det(\mathbf{D})}^{-j/d} \sup_{\bm{y} \in \mathbb{R}^d} \inf_{\bm{k}\in\mathbb{Z}^d} \abs{ \bm{y} - \bm{k} } \\ &= \pi \abs{\det(\mathbf{D})}^{-j/d} \sqrt{d}, \end{align*} and likewise \begin{equation*} q_j = \pi \abs{\det(\mathbf{D})}^{-j/d}. \end{equation*} Considering the distance function \begin{equation*} \text{dist}(\bm{0},X_j(m,\bm{\omega})) := \min_{\bm{x} \in X_j(m,\bm{\omega})} \abs{\bm{x}}, \end{equation*} we bound \eqref{eq:psi_sum} by considering two cases: \begin{enumerate} \item $\text{dist}(\bm{0},X_j(m,\bm{\omega})) \geq q_j/2$; \item $\text{dist}(\bm{0},X_j(m,\bm{\omega})) < q_j/2$. \end{enumerate} For Case 1, all points of the lattice $X_j(m,\bm{\omega})$ lie outside of the ball of radius $\omega_0$ centered at the origin. Therefore, \eqref{eq:psi_sum} can be reduced to \begin{equation}\label{eq:psi_sum_reduced} \left( \sum_{\bm{x}_{\bm{k}} \in X_j(m,\bm{\omega})} \abs{\widehat{L}(\bm{x}_{\bm{k}})}^{-2} \right)^{-1}. \end{equation} Applying Proposition \ref{pr:lattice_lower}, we can bound \eqref{eq:psi_sum_reduced} from above by a constant multiple of $h_j^{2r}=(\pi\sqrt{d})^{2r} \abs{\det(\mathbf{D})}^{-2rj/d}$, and applying Proposition \ref{pr:lattice_upper}, we bound \eqref{eq:psi_sum_reduced} from below by a constant multiple of $q_j^{2r}=\pi^{2r}\abs{\det(\mathbf{D})}^{-2rj/d}$. Importantly, the proportionality constants are independent of $j$. For Case 2, we must be more careful, as one of the lattice points lies close to the origin. However, for any fixed $\bm{\omega}$, there is at most one $m$ for which $\text{dist}(\bm{0},X_j(m,\bm{\omega})) < q_j/2$. Therefore, in this case, a sufficient lower bound for \eqref{eq:psi_sum} is $0$; however, the upper bound must match the one derived in Case 1. Let us further separate Case 2 into the cases \begin{enumerate}[2a.] \item $\widehat{L}$ takes the value $0$ at some point of the lattice $X_j(m,\bm{\omega})$ \item $\absf{\widehat{L}^{-1}}<\infty$ for every point of the lattice $X_j(m,\bm{\omega})$ \end{enumerate} In Case 2a, we see that \eqref{eq:psi_sum} is $0$. In Case 2b, we again reduce \eqref{eq:psi_sum} to \eqref{eq:psi_sum_reduced}, and we see that $0$ is a lower bound for \eqref{eq:psi_sum_reduced}. For the upper bound, we apply Proposition \ref{pr:lattice_lower}, and the bound coincides with the one obtained in Case 1. To finish the proof, we note that Lemmas \ref{lem:eig_values}, \ref{lem:rb}, and \ref{lem:rb_w} imply that the wavelets \begin{equation*} \left\{ \psi_{j+1}(\cdot-\beta)\right\}_{\beta \in \mathbf{D}^{j}\mathbb{Z}^d\backslash\mathbf{D}^{j+1}\mathbb{Z}^d } \end{equation*} form a Riesz basis at each level $j\leq j_0$. Furthermore, the bounds obtained here on $c(m,\bm{\omega})$ imply that the Riesz bounds are proportional to $\abs{\det(\mathbf{D})}^{(1/2-r/d)2j}$. Therefore, the collection \eqref{eq:rb} is a Riesz basis of $L_2(\mathbb{R}^d)$. \end{proof} Let us remark that in the proof of this theorem, we used the fact that the lattices corresponding to the matrix $\mathbf{D}$ scale uniformly in all directions. This allowed us to find upper and lower Riesz bounds that are independent of $j$. However, the Riesz bounds would depend on $j$ for general integer dilation matrices. \section{Decorrelation of Coefficients}\label{sec:decor} As was stated in the introduction, the primary reason for our construction is to promote a sparse wavelet representation. Our model is based on the assumption that the wavelet coefficients of a signal $s$ are computed by the $L_2$ inner product \begin{equation*} \left<s,\psi_j(\cdot-\mathbf{D}^j\bm{k})\right>. \end{equation*} Here, we should point out that, unless the wavelets form an orthogonal basis, reconstruction will be defined in terms of a dual basis. However, as our focus in this paper is the sparsity of the coefficients, we are content to work with the analysis component of the approximation and leave the synthesis component for future study. Now, considering our stochastic model, it is important to use wavelets that (nearly) decorrelate the signal within each scale, and one way to accomplish this goal is by modifying the underlying operator. Hence, given a spline-admissible pair, $\Lop$ and $\mathbf{D}$, we define a new spline-admissible pair, $\Lop_n$ and $\mathbf{D}$, by $\widehat{L}_{n}:=\widehat{L}^n$, and we shall see that as $n$ increases, the wavelet coefficients become decorrelated. This result follows from the fact that the $(\Lop_n)^*\Lop_n$-spline interpolants (appropriately scaled) converge to a sinc-type function, and it is motivated by the work of Aldroubi and Unser, which shows that a large family of spline-like interpolators converge to the ideal sinc interpolator \cite{unser94}. To state this result explicitly, we denote the generalized B-splines for $\Lop_n$ by $\widehat{\varphi}_{n,j}=\widehat{\varphi}_{ j}^n $. Therefore the $(\Lop_n)^*\Lop_n$-spline interpolants are given by \begin{align*} \widehat{\phi}_{n,j}(\bm{\omega}) &= \abs{\det(\mathbf{D})}^{j}\frac{\abs{\widehat{\varphi}_{n,j}(\bm{\omega})}^2}{\sum_{\bm{k} \in \mathbb{Z}^d} \abs{\widehat{\varphi}_{n,j}(\bm{\omega}+ 2\pi (\mathbf{D}^T)^{-j} \bm{k})}^2 } \\ &= \abs{\det(\mathbf{D})}^{j}\frac{\abs{\widehat{\varphi}_{j}(\bm{\omega})}^{2n}}{\sum_{\bm{k} \in \mathbb{Z}^d} \abs{\widehat{\varphi}_{j}(\bm{\omega}+ 2\pi (\mathbf{D}^T)^{-j} \bm{k})}^{2n} }, \end{align*} and we analogously define \begin{equation*} m_{n,j}(\bm{\omega}) = \frac{\abs{\widehat{\varphi}_{j}(\bm{\omega})}^{2n}}{\sum_{\bm{k} \in \mathbb{Z}^d} \abs{\widehat{\varphi}_{j}(\bm{\omega}+ 2\pi (\mathbf{D}^T)^{-j} \bm{k})}^{2n} }. \end{equation*} For any fundamental domain $\Omega_j$ of the lattice $2\pi(\mathbf{D}^T)^{-j}\mathbb{Z}^d$, let $\chi_{\Omega_j}$ denote the associated characteristic function. Proving decorrelation depends on showing that the functions $m_{n,j}$ converge almost everywhere to some characteristic function $\chi_{\Omega_j}$. This analysis is closely related to the convergence of cardinal series as studied in \cite{deboor86}. Our proof relies on the techniques used by Baxter to prove the convergence of the Lagrange functions associated with multiquadric functions \cite[Chapter 7]{baxter92}. The idea is to define disjoint sets covering $\mathbb{R}^d$. Each set has a single point in any given fundamental domain, and we analyze the convergence of $m_{n,j}$ on these sets. \begin{definition}\label{def:funddom} Let $\Omega_j$ be a fundamental domain of $2\pi(\mathbf{D}^T)^{-j}\mathbb{Z}^d$. For each $j\in \mathbb Z$ and for each $\bm{x}\in \Omega_j$, define the set \begin{equation*} E_{j,\bm{x}} := \left\{\bm{x}+2\pi (\mathbf{D}^T)^{-j} \bm{k}:\bm{k} \in \mathbb{Z}^d\right\}. \end{equation*} Since $\phi_{1,j}$ generates a Riesz basis, each set $E_{j,\bm{x}}$ has a finite number of elements $\bm{y}$ with $m_{1,j}(\bm{y})$ of maximal size. We define $F_j$ to be the set of $\bm{x}\in\Omega_j$ such that there is not a unique $\bm{y}\in E_{j,\bm{x}}$ where $m_{1,j}$ attains a maximum; i.e., $\bm{x}$ is in the complement of $F_j$ if there exists $\bm{y}\in E_{j,\bm{x}}$ such that \begin{equation*} m_{1,j}(\bm{y}) > m_{1,j}\left(\bm{y}+2\pi (\mathbf{D}^T)^{-j} \bm{k}\right) \end{equation*} for all $\bm{k}\neq \bm{0}$. \end{definition} \begin{lemma}\label{lem:decor} Let $\bm{x} \in \Omega_j\backslash F_j$, then for $ \bm{y}\in E_{j,\bm{x}} $ we have $m_{n,j}(\bm{y})\rightarrow 0$ if and only if $m_{1,j}(\bm{y})$ is not of maximal size over $E_{j,\bm{x}}$. Furthermore, if $m_{1,j}(\bm{y})$ is of maximal size, then $m_{n,j}(\bm{y})\rightarrow 1$. \end{lemma} \begin{proof} Fix $\bm{x}\in \Omega_j\backslash F_j$ and $y\in E_{j,\bm{x}}$. Notice that the periodicity of the denominator of $m_{1,j}$ implies that $m_{1,j}(\bm{y})$ is maximal iff $\abs{\widehat{\varphi}_{1,j}(\bm{y})}$ is maximal. Let us first suppose $m_{1,j}(\bm{y})$ is not maximal. If $\abs{\widehat{\varphi}_{1,j}(\bm{y})}=0$, the result is obvious. Otherwise, there is some $\bm{k}_0 \in \mathbb{Z}^d$ and $b<1$ such that \begin{equation*} \abs{\widehat{\varphi}_{1,j}(\bm{y})} \leq b \abs{\widehat{\varphi}_{1,j}(\bm{y}+2\pi (\mathbf{D}^T)^{-j}\bm{k}_0)}. \end{equation*} Therefore \begin{align*} \abs{\widehat{\varphi}_{n,j}(\bm{y})}^2 &\leq b^{2n} \abs{\widehat{\varphi}_{n,j}(\bm{y}+2\pi (\mathbf{D}^T)^{-j}\bm{k}_0)}^2 \\ &\leq b^{2n} \sum_{\bm{k} \in \mathbb{Z}^d} \abs{\widehat{\varphi}_{n,j}(\bm{y}+2\pi(\mathbf{D}^T)^{-j} \bm{k})} ^2 \end{align*} and the result follows. Next, suppose $m_{1,j}(\bm{y})$ is of maximal size. Since $\widehat{\varphi}_{1,j}$ has no periodic zeros, $\abs{\widehat{\varphi}_{1,j}(\bm{y})}\neq 0$. Therefore \begin{equation*} m_{n,j}(\bm{y}) = \frac{1}{B_{n,j}(\bm{y})} \end{equation*} with \begin{equation*} B_{n,j}(\bm{y}) = \sum_{\bm{k} \in \mathbb{Z}^d} \abs{ \frac{\widehat{\varphi}_{1,j}(\bm{y}+2\pi (\mathbf{D}^T)^{-j} \bm{k})}{\widehat{\varphi}_{1,j}(\bm{y})} }^{2n}. \end{equation*} Since $\abs{\widehat{\varphi}_{1,j}(\bm{y})}$ is of maximal size, all terms of the sum except one are less than $1$. In particular, $B_{n,j}$ will converge to $1$ as $n$ increases. \end{proof} \begin{lemma} Let $j \in \mathbb Z$, and let $\Omega_j$ be a fundamental domain. If the Lebesgue measure of $F_j$ is $0$, then \begin{equation*} \sum_{\bm{k} \in \mathbb{Z}^d} m_{n,j}(\cdot+2\pi (\mathbf{D}^T)^{-j}\bm{k})^2 \end{equation*} converges to $\chi_{\Omega_j}$ in $ L_1(\Omega_j)$ as $n\rightarrow \infty$. \end{lemma} \begin{proof} The sum is bounded above by $1$, so Lemma \ref{lem:decor} implies that it converges to $\chi_{\Omega_j}$ on the complement of $F_j$. Hence, we apply the dominated convergence theorem to obtain the result. \end{proof} With this theorem, we show how the wavelets corresponding to $\Lop_n$ decorrelate within scale as $n$ becomes large. The way we characterize decorrelation is in terms of the semi-inner products \begin{equation*} (f,g)_{n} := \int_{\mathbb{R}^d} \widehat{f} \widehat{g}^* \abs{\widehat{L}_{n}}^{-2}, \end{equation*} which are true inner products for the wavelets \begin{equation*} \psi_{n,j+1} = \Lop_{n}^*\phi_{n,j}. \end{equation*} \begin{theorem}\label{th:dc} Suppose the Lebesgue measure of $\cup_{j\in \mathbb{Z}} F_j$ is $0$, where $F_j$ is from Definition \ref{def:funddom}. Then as $n$ increases, the wavelet coefficients decorrelate in the following sense. For any $j\in\mathbb Z, \bm{k}\in \mathbb{Z}^d\backslash \{0\}$ we have \begin{equation*} (\psi_{n,j+1},\psi_{n,j+1}(\cdot- \mathbf{D}^{j}\bm{k} ))_n \rightarrow 0 \end{equation*} as $n \rightarrow \infty$. \end{theorem} \begin{proof} First, we express the inner product as an integral \begin{align*} (\psi_{n,j+1},\psi_{n,j+1}(\cdot- \mathbf{D}^{k}\bm{k}))_n &= \int_{\mathbb R^d} \widehat{\psi}_{n,j+1}(\bm{\omega}) (\psi_{n,j+1}(\cdot-\mathbf{D}^{j}\bm{k}))^\wedge(\bm{\omega})^* \abs{\widehat{L}_n}^{-2} \rmd\bm{\omega}\\ &= \abs{\det(\mathbf{D})}^{2j}\int_{\mathbb R^d} m_{n,j}(\bm{\omega})^2e^{- i\mathbf{D}^{j}\bm{k}\cdot\bm{\omega}}\rmd\bm{\omega}, \end{align*} and we periodize the integrand to get \begin{align*} \int_{\mathbb R^d} m_{n,j}(\bm{\omega})^2e^{- i\mathbf{D}^{j}\bm{k}\cdot\bm{\omega}}d\bm{\omega} &= \sum_{\bm{l}\in 2\pi (\mathbf{D}^T)^{-j}\mathbb{Z}^d} \int_{\Omega_j +\bm{l}} m_{n,j}(\bm{\omega})^2e^{- i\mathbf{D}^{j}\bm{k}\cdot\bm{\omega}} \rmd\bm{\omega} \\ &= \int_{\Omega_j} e^{- i\mathbf{D}^{j}\bm{k}\cdot\bm{\omega}}\sum_{\bm{l}\in 2\pi (\mathbf{D}^T)^{-j}\mathbb{Z}^d} m_{n,j}(\bm{\omega}-\bm{l})^2\rmd\bm{\omega}. \end{align*} The last expression converges to $0$ by the Lebesgue dominated convergence theorem. \end{proof} Let us now show how this result implies decorrelation of the wavelet coefficients. Recall that our model for a random signal $s$ is based on the stochastic differential equation $\Lop s =w$, where $w$ is a non-Gaussian white noise \cite{unser_part1} and the operator $\Lop$ is spline admissible. We denote the wavelet coefficients as \begin{align*} c_{n,j+1,\bm{k}} &:= \left<s,\psi_{n,j+1}(\cdot-\mathbf{D}^j\bm{k}) \right> \\ &= \left< w, \phi_{n,j}(\cdot-\mathbf{D}^j\bm{k}) \right>. \end{align*} Our stochastic model implies that the coefficients are random variables. Hence, for distinct $\bm{k}$ and $\bm{k}'$, the covariance between $c_{n,j+1,\bm{k}}$ and $c_{n,j+1,\bm{k}'}$ is determined by the expected value of their product: \begin{equation*} \mathbb{E}\{ c_{n,j+1,\bm{k}} c_{n,j+1,\bm{k}'} \} = \mathbb{E}\left\{ \left<w,\phi_{n,j}(\cdot-\mathbf{D}^j\bm{k}) \right>\left<w,\phi_{n,j}(\cdot-\mathbf{D}^j\bm{k}')\right> \right\}. \end{equation*} As long as the white noise $w$ has zero mean and finite second-order moments, the covariance satisfies \begin{align*} \mathbb{E}\{ c_{n,j+1,\bm{k}} c_{n,j+1,\bm{k}'} \} &= \left<\phi_{n,j}(\cdot-\mathbf{D}^j\bm{k}), \phi_{n,j}(\cdot-\mathbf{D}^j\bm{k}') \right> \\ &= (\psi_{n,j+1}(\cdot- \mathbf{D}^{j}\bm{k} ),\psi_{n,j+1}(\cdot- \mathbf{D}^{j}\bm{k}' ))_n \\ &\rightarrow 0, \end{align*} where the convergence follows from Theorem \ref{th:dc}. Therefore, when the standard deviations of $c_{n,j+1,\bm{k}}$ and $c_{n,j+1,\bm{k}'}$ are bounded below, the correlation between the coefficients converges to zero. \section{Discussion and Examples}\label{sec:conclusion} The formulation presented in this paper is quite general and accommodates many operators. In this section, we show how it relates to previous wavelet constructions, and we provide examples that are not covered by previous theories. \subsection{Connection to previous constructions}\label{sec:connection} Our operator-based wavelet construction can be viewed as a direct generalization of the cardinal spline wavelet construction. To see this, define the B-spline $N_1$ to be the characteristic function of the interval $[0,1]\subset \mathbb{R}$. Then for $m=2,3,\dots$, let the B-splines $N_m$ be defined by \begin{equation*} N_m(x) := \int_0^1 N_{m-1} (x-t) {\rm d} t. \end{equation*} The fundamental (cardinal) interpolatory spline $\phi_{2m}$ is then defined as the linear combination \begin{equation}\label{eq:interp_bspline} \phi_{2m} (x) := \sum_{k\in\mathbb{Z}} \alpha_{k,m} N_{2m} (x+m-k), \end{equation} satisfying the interpolation conditions: \begin{equation*} \phi_{2m}(k) = \delta_{k,0}, \quad k\in\mathbb{Z}, \end{equation*} where $\delta$ denotes the Kronecker delta function. In \cite{chui91}, the authors define the cardinal B-spline wavelets (relative to the scaling function $N_m$) as \begin{equation*} \psi_m(x) := \left(\frac{{\rm d}^m}{{\rm d} x^m}\phi_{2m} \right) (2x-1) \end{equation*} Within the context of our construction, the operator $\Lop$ is a constant multiple of the order $m$ derivative, and $\Lop^*\Lop$ is a constant multiple of the order $2m$ derivative. Therefore our $\Lop^*\Lop$ spline interpolants \eqref{eq:lagint} are equivalent to the $\phi_{2m}$ defined in \eqref{eq:interp_bspline}, so we obtain the same wavelet spaces. In particular, $\psi_m$ is a constant multiple of $\Lop^* \phi_{2m}$. A more general construction is given in \cite{micchelli91}. In that paper, the the authors allow for scaling functions $\varphi$ that are defined in the Fourier domain by $\widehat{\varphi} = T/q$, where $T$ is a trigonometric polynomial \begin{equation*} T(\bm{\omega}) := \sum_{\bm{k}\in \mathbb{Z}^d} c[\bm{k}] e^{-i \bm{k}\cdot \bm{\omega}}, \quad \bm{\omega} \in \mathbb{R}^d \end{equation*} and $q$ is a homogeneous polynomial \begin{equation*} q(\bm{\omega}) := \sum_{\abs{\bm{k}}=m} q_{\bm{k}} \bm{\omega}^{\bm{k}},\quad \bm{\omega} \in \mathbb{R}^d \end{equation*} of degree $m$ with $m>d$. Here, $q$ is also required to be elliptic; i.e., $q$ can only be zero at the origin. The authors then define the Lagrange function $\phi$ by \begin{equation*} \widehat{\phi} = \frac{\abs{\widehat{\varphi}}^2}{\sum_{\bm{k}\in\mathbb{Z}^d} \abs{\widehat{\varphi}(\cdot+2\pi \bm{k}) }^2 }, \end{equation*} and they define an elliptic spline wavelet $\psi_0$ as \begin{equation*} \widehat{\psi}_0 = 2^{-d} q^* \widehat{\phi}(2^{-1}\cdot). \end{equation*} Thus, we can see that our construction is also a generalization of the elliptic spline wavelet construction, the primary extension being that we allow for a broader class of operators. In fact, both of the prior constructions use scaling functions associated with operators that have homogeneous symbols. This special case has the property that the multiresolution spaces can be generated by dilation. \begin{proposition}\label{pr:homogen} Let $\Lop$ be an admissible Fourier multiplier operator whose symbol $\widehat{L}$ is positive (except at the origin), continuous, and homogeneous of order $\alpha>d/2$; i.e., \begin{equation*} \widehat{L}(a\bm{\omega}) = a^\alpha\widehat{L}(\bm{\omega}), \quad \text{for}\quad a>0. \end{equation*} Further assume that there is a localization operator $\Lop_{{\rm d},0}$ (of the form described in Definition \ref{def:spadm}) such that the generalized B-spline $\varphi_0$, defined by \begin{equation*} \widehat{\varphi}_0 (\bm{\omega}) := \frac{\widehat{L}_{{\rm d},0}(\bm{\omega})}{\widehat{L}(\bm{\omega})}, \end{equation*} satisfies the Riesz basis condition \begin{equation*} 0 < A \leq \sum_{\bm{k}\in\mathbb{Z}^d}\abs{\widehat\varphi_0(\bm{\omega}+2\pi \bm{k})}^2 \leq B < \infty, \end{equation*} for some $A$ and $B$ in $\mathbb{R}$. Then the pair $\Lop, \mathbf{D}=2\mathbf{I}$ is spline admissible of order $\alpha$. \end{proposition} \begin{proof} The first two conditions of Definition \ref{def:spadm} are automatically satisfied. For the third condition, let $C_{\Lop}>0$ be a constant satisfying \begin{equation*} C_{\Lop} \abs{\widehat{L}(\bm{\omega})}^2 \geq 1 \end{equation*} on the unit sphere $\mathbb{S}^{d-1}\subset\mathbb{R}^d$. Then for $\abs{\bm{\omega}}>0$, homogeneity of $\widehat{L}$ implies \begin{align*} C_{\Lop} \left(1+ \abs{\widehat{L}(\bm{\omega})}^2 \right) &\geq C_{\Lop} \abs{\widehat{L} \left( \abs{\bm{\omega}} \frac{\bm{\omega}}{\abs{\bm{\omega}}} \right)}^2 \\ & \geq \abs{\bm{\omega}}^{2\alpha}. \end{align*} For the fourth property, we let $\widehat{L}_{{\rm d},0}(\bm{\omega})= \sum_{\bm{k}\in \mathbb{Z}^d}p[\bm{k}]e^{i\bm{\omega}\cdot \bm{k}}$ and define \begin{equation*} \widehat{L}_{{\rm d},j}(\bm{\omega}) := 2^{-j\alpha}\sum_{\bm{k}\in \mathbb{Z}^d}p[\bm{k}]e^{i\bm{\omega}\cdot 2^j\bm{k}}. \end{equation*} The generalized B-splines $\varphi_j$ will then satisfy \begin{align*} \widehat{\varphi}_j(\bm{\omega}) &= \widehat{L}_{{\rm d},j}(\bm{\omega}) \widehat{L}(\bm{\omega})^{-1}\\ &= 2^{-j\alpha}\widehat{L}_{{\rm d},0}(2^j\bm{\omega}) \widehat{L}(\bm{\omega})^{-1}\\ &= \widehat{L}_{{\rm d},0}(2^j\bm{\omega}) \widehat{L}(2^{j}\bm{\omega})^{-1}\\ &= \widehat{\varphi}_0(2^j\bm{\omega}). \end{align*} The Riesz basis property can then be verified, since \begin{equation*} \sum_{\bm{k}\in\mathbb{Z}^d}\abs{\widehat\varphi_j(\bm{\omega}+2\pi 2^{-j}\bm{k})}^2 = \sum_{\bm{k}\in\mathbb{Z}^d}\abs{\widehat\varphi_0(2^{j}\bm{\omega}+2\pi \bm{k})}^2. \end{equation*} \end{proof} In summary, our wavelet construction generalizes these known constructions for homogeneous Fourier multiplier operators, and it accommodates the more complex setting of non-homogeneous operators. \subsection{Mat\'ern and Laplace operator examples} \label{sec:examples_lap} The $d$-dimensional Mat\'ern operator is not homogeneous, so it provides an example that is not included in traditional wavelet constructions. Its symbol is $\widehat{L}_\nu(\bm{\omega})=(\abs{\bm{\omega}}^2+1)^{\nu/2}$, with the parameter $\nu > d/2$. As $\widehat{L}_\nu(\bm{\omega})^{-1}$ satisfies the Riesz basis condition, no localization operator is needed. Therefore, the operator $\Lop_\nu$ is spline-admissible of order $\nu$ for any admissible subsampling matrix $\mathbf{D}$. Next, consider the iterated Laplacian operator with symbol $\widehat{L}=\abs{\bm{\omega}}^{2m}$, where $m>d/4$ is an integer, and let $\mathbf{D}=2\mathbf{I}$. Localization operators $\Lop_{{\rm d},j}$ can be constructed as in \cite{chui92}, and all of the conditions of Definition \ref{def:spadm} are satisfied. Note that each of these operators satisfies the growth condition of Theorem \ref{th:rb2}, so the corresponding wavelet spaces may be used to construct Riesz bases of $L_2(\mathbb{R}^d)$. \subsection{Construction of non-standard localization operators} \label{sec:examples_helm} Here, we consider the Helmholtz operator $\Lop$ and construct corresponding localization operators $\Lop_{\rmd,j}$. While we focus on this particular operator, the presented method is sufficient to be applied more generally. The Helmholtz operator is defined by its symbol $\widehat{L}(\omega)=1/4-\abs{\bm{\omega}}^2$. The wavelets corresponding to $\Lop$ could potentially be applied in optics, as the Helmholtz equation, \begin{equation*} \Delta u + \lambda u = f, \end{equation*} is a reduced form of the wave equation \cite[Chapter 5]{duffy01}. In what follows, we show that this is a spline-admissible operator for the scaling matrix $\mathbf{D}=2\mathbf{I}$ on $\mathbb{R}^2$. However, since the wavelets $\psi_{j+1}=\Lop^*\phi_j$ do not form a Riesz basis for the coarse-scale wavelet spaces, we only consider $j\leq 0$. Our construction of localization operators $\Lop_{\rmd, j}$ is based on the fact that sufficiently smooth functions have absolutely convergent Fourier series \cite[Theorem 3.2.9]{grafakos08}. This implies that we can define $\Lop_{\rmd, j}$ by constructing smooth, periodic functions $\widehat{L}_{\rmd, j}$ that are asymptotically equivalent to $\widehat{L}$ at its zero set. In fact, we define $\widehat{L}_{\rmd, j}$ to be equal to $\widehat{L}$ near its zeros. Notice that $\widehat{L}$ is zero on the circle of radius $1/4$ centered at the origin, and it is smooth in a neighborhood of this circle. Therefore, in the case $j=0$, we choose $\epsilon>0$ sufficiently small and define $\widehat{L}_{\rmd,0}$ to be a function satisfying: \begin{enumerate} \item $\widehat{L}_{\rmd,0}(\bm{\omega})=\widehat{L}(\bm{\omega})$ for $\abs{\abs{\bm{\omega}}-1/4}<\epsilon$; \item $\widehat{L}_{\rmd,0}(\bm{\omega})$ is constant for $\abs{\abs{\bm{\omega}}-1/4}>3\epsilon$ and $\omega \in [-\pi,\pi]^2$; \item $\widehat{L}_{\rmd,0}(\bm{\omega})$ is periodic with respect to the lattice $2\pi \mathbb{Z}^2$. \end{enumerate} Such a function can be constructed using a smooth partition of unity $\{f_n\}_{n=1}^N$ on the torus, where each $f_n$ is supported on a ball of radius $\epsilon$. Here, we require each $f_n$ to be positive, and the partition of unity condition means that \begin{equation*} \sum_{n=1}^N f_n(\bm{\omega}) = 1. \end{equation*} We partition the index set $\{1,\dots,N\}$ into the three subsets $\Lambda_1,\Lambda_2,\Lambda_3$ as follows: \begin{enumerate} \item If the support of $f_n$ has a non-empty intersection with the annulus $\abs{\abs{\bm{\omega}}-1/4}\leq \epsilon$, then $n\in \Lambda_1$; \item Else if the support of $f_n$ lies in the ball of radius $1/4-\epsilon$ centered at the origin, then $n\in \Lambda_2$; \item Else $n\in \Lambda_3$. \end{enumerate} We now define the periodic function \begin{equation*} \widehat{L}_{\rmd,0} := \sum_{n\in \Lambda_1} f_n \widehat{L} + \sum_{n\in \Lambda_2} f_n - \sum_{n\in \Lambda_3} f_n, \end{equation*} on $[-\pi,\pi]^2$, and it can be verified that this function has the required properties. Using a similar approach, we can define $\widehat{L}_{\rmd, j}$ for $j<0$, and the conditions of spline admissibility can be verified. Since the Helmholtz operator satisfies the conditions of Theorem \ref{th:rb2}, the resulting wavelet system is a Riesz basis of $L_2(\mathbb{R}^d)$. In conclusion, we have constructed localization operators (and hence generalized B-splines) for the Helmholtz operator. Furthermore, the presented method applies in greater generality to operators whose symbols are smooth near their zero sets.
\section{Introduction} Over the last three decades, cosmology has made tremendous progress, culminating in the so-called ``Standard Model of Cosmology''. The two main components of the Standard Model are a mysterious dark energy, leading to a late-time accelerated expansion, and a dark matter component making up roughly 25\% of the total matter-energy budget of the Universe. The evolution of structure in the Universe from the earliest accessible times to today is successfully described by a theory based on the gravitational instability -- the distribution of galaxies in the Universe, for example, is remarkably well reproduced by this paradigm. Clusters and groups of galaxies are major building blocks of the large-scale structure and measurements of their abundance provide a powerful cosmological probe. Large, gravity-only $N$-body simulations have been remarkably successful in providing a consistent picture of the formation of the large-scale structure from the very early, small Gaussian density fluctuations to the halos, voids, and filaments we observe today. A surprising discovery from these simulations~\citep{nfw1,nfw2} was that the dark matter-dominated halos -- over a wide mass range typical of dwarf galaxies to massive clusters -- share a basically universal density profile. In detail, it was shown that the spherically averaged density profile of relaxed halos formed in simulations can be described by what is now commonly known as the NFW (Navarro-Frenk-White) profile. The NFW profile is described by two parameters, the normalization and the characteristic scale radius of the halo or equivalently its (dimensionless) concentration. Aside from the distribution of halo masses, halo profiles are also of considerable interest. The profiles can be measured directly for individual massive halos by a variety of observational methods, or inferred indirectly for less massive halos using statistical lensing probes. Halo profiles are also a key input in halo occupation distribution (HOD) modeling of the distribution of galaxies. NFW profiles (or minor variants thereof) are consistent with current observations~\citep{bhattacharya11_b} and, as the observations continue to improve, a corresponding improvement in theoretical predictions for (NFW) halo concentrations as a function of cosmological parameters is needed. As scatter in the $c-M$ relation is considerable, in principle, this would encompass knowing the actual distribution of halo concentrations as a function of halo mass. Quantitative predictions for the $c-M$ relation from a first principles analytic approach are difficult to obtain, due to the highly nonlinear dynamics involved in the formation of halos. Accurate predictions can only be obtained from computationally expensive, high-resolution simulations. These simulations need to cover large volumes in order to yield good statistics, especially in the cluster mass regime, as well as high force resolution to reliably resolve the halo profiles. In recent years, the focus has therefore been on generating predictions for one cosmology around the best-fit WMAP (Wilkinson Microwave Anisotropy Probe) results of that time~(e.g., \citealt{duffy08}, \citealt{bhattacharya11_b}, \citealt{prada12}). The fitting functions so generated cannot be extended beyond the cosmological model they have been tuned for. Heuristic models that aim to extend this reach, e.g. those by \cite{bullock} and \cite{eke01} and improvements thereof (\citealt{maccio08}) do not lead to the desired accuracy, as discussed in \cite{duffy08}. Other discussions of this issue can be found in ~\cite{gao07}, \cite{hayashi07}, and \cite{zhao09}. In order to overcome the many shortcomings of fitting functions as a general approach in cosmology, we have recently developed the ``Cosmic Calibration Framework'' (CCF) to provide accurate prediction schemes for cosmological observables~\citep{HHHN,HHHNW}. The aim of the CCF is to build codes that act as very fast -- basically instantaneous -- prediction tools for large scale structure observables such as the nonlinear power spectrum~\citep{coyote1,coyote2,coyote3}, mass functions for different halo definitions, or the concentration-mass relation -- as discussed here. Predicting these observables requires running a number of high-performance simulations to reliably resolve the nonlinear regime of structure formation. The CCF provides a powerful way to build precision prediction tools from a limited number of computationally expensive simulations. At the heart of the CCF lies a sophisticated sampling scheme that provides an optimal sampling strategy for the cosmological models to be simulated (we use orthogonal array-based Latin hypercube as well as symmetric Latin hypercube designs; an introduction to the general sampling strategy is provided in~\citealt{santner03}), an optimal representation to translate the measurements from the simulations into functions that can be easily interpolated (a principal component basis turns out to be an efficient representation), and finally a very accurate interpolation scheme (our choice here is Gaussian process modeling). The CCF was first introduced in~\cite{HHHN} and a more detailed description and examples are provided in~\cite{HHHNW}. In a series of three papers (Coyote Universe I-III) we developed an emulator for the matter power spectrum for a five dimensional parameter space covering $\theta=\{\omega_b, \omega_m, n_s, w, \sigma_8\}$. This emulator provides predictions for the power spectrum for $w$CDM cosmologies out to $k\sim 1$~Mpc$^{-1}$ at the 1\% accuracy level for a redshift range of $0\le z\le 1$. In~\cite{SKHHHN} the work was extended to derive an approximate statistical model for the sample variance distribution of the nonlinear matter power spectrum. \cite{eifler} used the emulator to generate a weak lensing prediction code to calculate various second-order cosmic shear statistics, e.g., shear power spectrum, shear-shear correlation function, ring statistics and Complete Orthogonal Set of EB-mode Integrals (COSEBIs). The focus of this paper is the development of an emulator for the concentration-mass relation for $w$CDM cosmologies. We use the same base set of simulations as in \cite{coyote3}, consisting of 37 cosmological models and a single 1300~Mpc volume, high-resolution simulation for each model. This simulation set is augmented here with a set of new, higher resolution simulations. These simulations cover smaller volumes (a 360~Mpc and a 180~Mpc simulation for each model) to obtain good statistics over a large range of halo masses. For each model we measure the best-fit concentration-mass ($c-M$) relation, assuming a simple power law form. The fits lay the foundation for building the emulator that provides predictions for the $c-M$ relation within the $w$CDM parameter space covered by the original simulations. In redshift, the emulator covers the range between $z=0$ and $z=1$. We provide a fast code that delivers the mean $c-M$ relations for $w$CDM cosmologies to good accuracy\footnote{{\tt http://www.hep.anl.gov/cosmology/CosmicEmu}}. As is well-known, the $c-M$ relation has considerable scatter and, in principle, it is not obvious that this scatter should have a simple form, and what its cosmological dependence might be. However, as discussed in~\cite{bhattacharya11_b}, the scatter has a simple Gaussian form in $w$CDM models, and moreover, even though the mean $c-M$ relation is clearly cosmology-dependent, as is the associated concentration variance, $\sigma_c^2(M)$, the ratio of $\sigma_c(M)$ to the mean concentration is close to $1/3$, independent of cosmology, mass, or redshift. This means that once an emulator for the $c-M$ relation is in hand, the concentration standard deviation is given automatically by a simple relation. The paper is organized as follows. After a brief outline of the halo concentration measurements from the simulations, we describe the cosmological model space and the simulation suite used to build the emulator. In Section~\ref{sec:models} we also discuss the generation of the smooth prediction for the concentration-mass relation for each model that underlies the interpolation scheme for building the emulator. We give a brief description on how to build the emulator in Section~\ref{sec:emu} and show some examples from the working emulator and test results verifying its accuracy. We also compare our results to currently used fitting formulae and investigate the cosmology dependence of the $c-M$ relation in some detail. Finally, we provide a conclusion and outlook in Section~\ref{sec:conc}. \section{Concentration-Mass Relation} \label{sec:cm} We study the concentration-mass relation in the regime of bright galaxies to clusters of galaxies, spanning halo mass ranges between $2\cdot 10^{12}$M$_\odot$ to $10^{15}$M$_\odot$, while varying $w$CDM cosmological parameters. A detailed description on how to measure halo concentrations from simulations and a discussion of possible systematics is given in~\cite{bhattacharya11_b}. We follow the same approach in this paper and give here a brief summary of the main steps in measuring the $c-M$ relation in our simulations. As a first step, we identify halos using a fast parallel friends-of-friends (FOF) finder \citep{woodring11} with linking length $b=0.2$. Once a halo is found, we define its center via a density maximum criteria -- the location of the particle with the maximum number of neighbors. This definition of the halo center is very close to that given by the halo's potential minimum. Given a halo center, we grow spheres around it and compute the mass in radial bins. Note that even though an FOF finder is used, the actual halo mass is defined by a spherical overdensity method, consistent with what is done in observations. (For discussions on halo mass, see, e.g., \citealt{white01}, \citealt{lukic09}, and \citealt{more11}). The NFW form for the spherically averaged halo profile is a function of two parameters, one of which is constrained by the halo mass. Here we fit the mass profile using both total halo mass and concentration as free variables. Although the mass could be measured independently of the concentration, the joint analysis is potentially less sensitive to fitting bias. We write the NFW profile as \begin{equation} \rho(r)= \frac{\delta\rho_{\rm{crit}}}{(r/r_s)(1+r/r_s)^2}, \label{eq:nfw} \end{equation} where $\delta$ is a characteristic dimensionless density, and $r_s$ is the scale radius of the NFW profile. The concentration of a halo is defined as $c_{\Delta}=r_{\Delta}/r_s$, where $\Delta$ is the overdensity with respect to the {\it critical density} of the Universe, $\rho_{\rm{crit}}=3H^2/8\pi G$, and $r_{\Delta}$ is the radius at which the enclosed mass, $M_{\Delta}$, equals the volume of the sphere times the density $\Delta \rho_{\rm{crit}}$. We compute concentrations corresponding to $\Delta=200$, corresponding in turn to $c_{200}= R_{200}/r_s$. \begin{table*} \begin{center} \caption{The parameters for the 37+1 models which define the sample space. See text for further details. \label{tab:basic}} \vspace{-0.3cm} \begin{tabular}{ccccccc|ccccccc} \# & $\omega_m$ & $\omega_b$ & $n_s$ & $-w$ & $\sigma_8$ & $h$ & \# & $\omega_m$ & $\omega_b$ & $n_s$ & $-w$ & $\sigma_8$ & $h$ \\ \hline M000 & 0.1296 & 0.0224 & 0.9700 & 1.000 & 0.8000 & 0.7200 & M019 & 0.1279 & 0.0232 & 0.8629 & 1.184 & 0.6159 & 0.8120 \\ M001 & 0.1539 & 0.0231 & 0.9468 & 0.816 & 0.8161 & 0.5977 & M020 & 0.1290 & 0.0220 & 1.0242 & 0.797 & 0.7972 & 0.6442 \\ M002 & 0.1460 & 0.0227 & 0.8952 & 0.758 & 0.8548 & 0.5970 & M021 & 0.1335 & 0.0221 & 1.0371 & 1.165 & 0.6563 & 0.7601 \\ M003 & 0.1324 & 0.0235 & 0.9984 & 0.874 & 0.8484 & 0.6763 & M022 & 0.1505 & 0.0225 & 1.0500 & 1.107 & 0.7678 & 0.6736 \\ M004 & 0.1381 & 0.0227 & 0.9339 & 1.087 & 0.7000 & 0.7204 & M023 & 0.1211 & 0.0220 & 0.9016 & 1.261 & 0.6664 & 0.8694 \\ M005 & 0.1358 & 0.0216 & 0.9726 & 1.242 & 0.8226 & 0.7669 & M024 & 0.1302 & 0.0226 & 0.9532 & 1.300 & 0.6644 & 0.8380 \\ M006 & 0.1516 & 0.0229 & 0.9145 & 1.223 & 0.6705 & 0.7040 & M025 & 0.1494 & 0.0217 & 1.0113 & 0.719 & 0.7398 & 0.5724 \\ M007 & 0.1268 & 0.0223 & 0.9210 & 0.700 & 0.7474 & 0.6189 & M026 & 0.1347 & 0.0232 & 0.9081 & 0.952 & 0.7995 & 0.6931 \\ M008 & 0.1448 & 0.0223 & 0.9855 & 1.203 & 0.8090 & 0.7218 & M027 & 0.1369 & 0.0224 & 0.8500 & 0.836 & 0.7111 & 0.6387 \\ M009 & 0.1392 & 0.0234 & 0.9790 & 0.739 & 0.6692 & 0.6127 & M028 & 0.1527 & 0.0222 & 0.8694 & 0.932 & 0.8068 & 0.6189 \\ M010 & 0.1403 & 0.0218 & 0.8565 & 0.990 & 0.7556 & 0.6695 & M029 & 0.1256 & 0.0228 & 1.0435 & 0.913 & 0.7087 & 0.7067 \\ M011 & 0.1437 & 0.0234 & 0.8823 & 1.126 & 0.7276 & 0.7177 & M030 & 0.1234 & 0.0230 & 0.8758 & 0.777 & 0.6739 & 0.6626 \\ M012 & 0.1223 & 0.0225 & 1.0048 & 0.971 & 0.6271 & 0.7396 & M031 & 0.1550 & 0.0219 & 0.9919 & 1.068 & 0.7041 & 0.6394 \\ M013 & 0.1482 & 0.0221 & 0.9597 & 0.855 & 0.6508 & 0.6107 & M032 & 0.1200 & 0.0229 & 0.9661 & 1.048 & 0.7556 & 0.7901 \\ M014 & 0.1471 & 0.0233 & 1.0306 & 1.010 & 0.7075 & 0.6688 & M033 & 0.1399 & 0.0225 & 1.0407 & 1.147 & 0.8645 & 0.7286 \\ M015 & 0.1415 & 0.0230 & 1.0177 & 1.281 & 0.7692 & 0.7737 & M034 & 0.1497 & 0.0227 & 0.9239 & 1.000 & 0.8734 & 0.6510 \\ M016 & 0.1245 & 0.0218 & 0.9403 & 1.145 & 0.7437 & 0.7929 & M035 & 0.1485 & 0.0221 & 0.9604 & 0.853 & 0.8822 & 0.6100 \\ M017 & 0.1426 & 0.0215 & 0.9274 & 0.893 & 0.6865 & 0.6305 & M036 & 0.1216 & 0.0233 & 0.9387 & 0.706 & 0.8911 & 0.6421 \\ M018 & 0.1313 & 0.0216 & 0.8887 & 1.029 & 0.6440 & 0.7136 & M037 & 0.1495 & 0.0228 & 1.0233 & 1.294 & 0.9000 & 0.7313 \end{tabular} \end{center} \end{table*} The mass enclosed within a radius $r$ for an NFW halo is given by \begin{equation} M(<r)= m(c_\Delta r/R_{200})/m(c_{200})M_{200}, \label{eq:nfwmass} \end{equation} where $m(y)=\ln(1+y)-y/(1+y)$. The mass in a radial bin is then \begin{equation} M_i=M(<r_i)-M(<r_{i-1}). \label{eq:nfwmassbin} \end{equation} We then fit Eq.~\ref{eq:nfwmassbin} to the mass contained in the radial bins of each halo, by minimizing the associated value of $\chi^2$ as \begin{equation} \chi^2= \sum_i \frac{(M_i^{sim}-M_i)^2}{(M_i^{sim})^2/n_i}, \label{eq:fit} \end{equation} where the sum is over the radial bins, $n_i$ is the number of particles in a radial bin, $M_i^{sim}$ is the mass in bin $i$ calculated from the simulations and $M_i$ is the mass calculated assuming the NFW profile. The advantage of fitting mass in radial bins rather than the density is that the bin center does not have to be specified. Note that we explicitly account for the finite number of particles in a bin. This leads to a slightly larger error in the profile fitting but minimizes any possible bias due to the finite number of particles, especially near the halo center. We fit for two parameters -- the normalization of the profile and the concentration. Halo profiles are fitted in the radial range of approximately $(0.1-1)R_{vir}$. This choice is motivated partly by the observations of concentrations that typically exclude the central region of clusters (e.g., observations by \citealt{oguri11}). More significantly, however, this excludes the central core which is sensitive to the effects of baryonic physics and numerical errors arising from limitations in both mass and force resolution. \cite{duffy10} have shown that, at $r<0.1R_{vir}$, cluster halo profiles are potentially sensitive to the impact of baryons with the profiles being affected at $r=0.05 R_{vir}$ by as much as a factor of 2. The $c-M$ relation is calculated by weighing the individual concentrations by the halo mass, \begin{equation} c(M)= \frac{\sum_i c_i M_i}{\sum_i M_i}, \label{eq:mean_c} \end{equation} where the sum is over the number, $N_i$, of the halos in a mass bin. The mass of the bin is given by \begin{equation} M= \sum_i M_i/N_i. \label{eq:m} \end{equation} The error on $c(M)$ is the mass-weighted error on the individual fits plus the Poisson error due to the finite number of halos in an individual bin added in quadrature, \begin{equation} \Delta c(M)= \sqrt {\left (\frac{\sum_i \Delta c_i M_i}{\sum_i M_i}\right )^2+ \frac{c^2(M)}{N_i}}, \label{eq:errmean} \end{equation} where $\Delta c_i$ is the individual concentration error for each halo. The first term dominates towards the lower mass end where the individual halos have smaller number of particles and the second term dominates towards the higher mass end, where there are fewer halos to average over. \section{Cosmological Models and Simulation Sets} \label{sec:models} \begin{figure*}[t] \centerline{ \includegraphics[width=7.5in]{cmmodels.eps}} \caption{\label{cm-rel}Concentration-mass relations for 37 $w$CDM cosmologies. The blue points show the measurements from the three simulations per model while the red lines show the best-fit power law for each measurement. In each subplot we show the results for $z=0$ and $z=1$ (upper and lower curve respectively). We find the best-fit power law separately for both redshifts. Models with low values for $\sigma_8$ in general also exhibit lower $c-M$ relations (e.g. M012, M018, M019). The fits shown here are the foundation for building the emulator described in Section~\ref{sec:emu}. } \end{figure*} We now describe the cosmological model space covered by our prediction scheme and the simulations used to construct it. The emulator is based on 37 cosmological models spanning the class of $w$CDM cosmologies. We allow for variations of the following five parameters: \begin{equation} \theta=\{\omega_b, \omega_m, n_s, w, \sigma_8\}. \end{equation} The 37 models are chosen to lie within the ranges: \begin{equation} \begin{array}{c} 0.0215 < \omega_b < 0.0235, \\ 0.120 < \omega_m < 0.155, \\ 0.85 < n_s < 1.05, \\ -1.30 < w <-0.70, \\ 0.616 < \sigma_8 < 0.9, \end{array} \label{priors} \end{equation} which are picked based on current constraints from CMB measurements~\citep{wmap7}. Following the approach in~\cite{coyote3} we lock the value of the Hubble parameter $h$ to the best-fit value for each model, given the measurement of the distance to the surface of last scattering. The values for $h$ then range from $0.55 < h < 0.85$. In addition to the 37 models, we run one $\Lambda$CDM model (M000 in Table~\ref{tab:basic}) which is not used to build the emulator. Instead we use this model as a control for testing the accuracy of the emulator. All 37+1 models are specified in detail in Table~\ref{tab:basic}. The specific model selection process is described at length in \cite{coyote2}. In summary, it is based on Symmetric Latin Hypercube (SLH) sampling~\citep{slh}; this sampling strategy provides a scheme that guarantees good coverage of the parameter hypercube. In our specific case we choose an SLH design that has good space filling properties in the case of two-dimensional projections in parameter space. In other words, if any two parameters are displayed in a plane, the plane will be well covered by simulation points. \cite{coyote2} provide an extensive discussion regarding optimal design choices and we refer the interested reader to that paper. \begin{table} \begin{center} \caption{Box sizes, particle numbers, and mass resolution. \label{tab:nest}} \begin{tabular}{ccccc} Length $[$Mpc$]$ & $N_p^3$ & Force res. [kpc] & $m_p$ $[$M$_\odot]$ \\ \hline 1300 & $1024^3$ & 50 & $5.7\cdot10^{11}\omega_m $ \\ 365 & $512^3$ & 10 & $1.0\cdot 10^{11}\omega_m $ \\ 180 & $512^3$ & 10 & $1.2\cdot 10^{10}\omega_m $ \\ \end{tabular} \end{center} \end{table} The emulator developed here is valid between $0<z<1$ and covers a halo mass range from $2\cdot 10^{12}$M$_\odot$ to $10^{15}$M$_\odot$. We use different box sizes to cover different mass ranges with sufficient statistics. A summary of the different simulation sizes is given in Table~\ref{tab:nest}. All simulations were carried out with the TreePM code {\sc GADGET-2}~\citep{gadget2}. In previous work \citep{bhattacharya11_b}, we have shown that results from {\sc GADGET-2} simulations and those with HACC \citep{habib09,pope10} produce completely consistent results. Results from a recent cluster re-simulation campaign \citep{wu12} are also in good agreement with those of \cite{bhattacharya11_b}. One set of simulations is from the original Coyote Universe suite as described in~\cite{coyote3}. This set of runs evolves 1024$^3$ particles in (1300~Mpc)$^3$ volumes. In addition, we run one realization each per model with 512$^3$ particles with a 10~kpc force resolution in a 365~Mpc box and a 180~Mpc box. A summary of the simulation sets including force and mass resolution is given in Table~\ref{tab:nest}. We combine the simulation results from the three boxes for each model to obtain measurements spanning the desired mass range. While some models (in particular those with high values of $\sigma_8$) have clusters at even higher masses, the statistics beyond 10$^{15}$M$_\odot$ are insufficient and we exclude those measurements. This is also done in order to avoid extrapolations for models where no data points at high masses exist. In order to build an emulator, for each model we have to provide a prediction for the $c-M$ relation for the same mass range. This ensures that we can provide a consistent set of measurements for the final interpolation process between different models. From the simulation results, we determine for each of the 37+1 cosmologies the best-fit $c-M$ relation by simply finding the best-fit power law for each model at two redshifts, $z=0$ and $z=1$. The results for the 37 models underlying the emulator are shown in Fig.~\ref{cm-rel}. The blue points show the simulation results while the red curves show the best-fit power law for each model. The upper curves in each plot are obtained at redshift $z=0$ and the lower curves at $z=1$. The concentration values range between $c\sim 2$ and $c\sim 8$. As expected, we find that models with low values of $\sigma_8$ (e.g., M012, M018, M019 with $\sigma_8<0.65$) have depressed $c-M$ relations. We will return to the cosmology dependence of the $c-M$ relation in Section~\ref{subsec:testing} after constructing the emulator, which will allow us to carry out a comprehensive sensitivity analysis. We reiterate that the fits shown in Fig.~\ref{cm-rel} are the basis for building the emulator; this procedure is discussed in the next section. \begin{center} \begin{figure}[t] \includegraphics[width=\linewidth]{sigmac.eps} \caption{Ratio of the standard deviation of the concentration to the mean concentration as a function of mass for all 37 cosmologies at $z=0$. While both the standard deviation and the mean concentration are functions of cosmology and redshift, their ratio is essentially invariant, and is approximately $1/3$. The distribution of the concentration around the mean is well-fit by a Gaussian distribution~\citep{bhattacharya11_b}.} \label{fig:var} \end{figure} \end{center} Finally we turn to a discussion of the intrinsic scatter in the $c-M$ relation. As mentioned earlier, the distribution of concentrations at any given halo mass is Gaussian, and the ratio of $\sigma_c(M)$ to the mean concentration is an approximate invariant for $w$CDM models, with a value of $\sim 1/3$~\citep{bhattacharya11_b}, independent of redshift and halo mass. This behavior is exhibited in Fig.~\ref{fig:var} where the ratio is computed for all 37 cosmologies as a function of halo mass, at $z=0$. Thus, given the $c-M$ relation from the emulator, the standard deviation at each mass bin can be trivially estimated by multiplying the returned concentration value by $1/3$. \section{Emulator for the Concentration-Mass Relation} \label{sec:emu} \subsection{Building the emulator}\label{subsec:building} In this section, we briefly outline the process for building the $c-M$ emulator. We follow the procedure explained in~\cite{coyote2} and refer the reader to this paper for more complete details. The focus of~\cite{coyote2} was on modeling the matter power spectrum rather than the $c-M$ relation, however, the process is essentially unchanged. Starting with the design of 37 models given in Table~\ref{tab:basic}, we measure the $c-M$ relation for each cosmology at $z=0$ and $z=1$ and fit these with a power law as described in Section~\ref{sec:models} to obtain a smooth functional form. First, for every mass bin, the global mean value is subtracted, and then via a simple rescaling, the concentrations are normalized to have unit variance. This produces a zero-mean, unit variance dataset spanning the 37 cosmologies. To reduce the dimensionality of the problem, these normalized functions are then decomposed into principal component (PC) basis functions and only the most significant components are kept. The idea is to apply the interpolation method of choice (Gaussian process modeling in our case) to the coefficients of the basis functions, rather than to the raw data itself (see~\citealt{coyote2} for details). Figure~\ref{fig:PC} shows that we only need three PCs to successfully capture the behavior of the $c-M$ relation since the shape of the relationship remains fairly simple across this set of cosmologies. As explained above, the emulator actually returns the weight on each PC basis function and these can be combined together to give the new $c-M$ relation. We model the error in the projection to the PC basis with an additional hyperparameter $\lambda_p$ that can be tuned to represent the level of noise in the data. \begin{center} \begin{figure}[t] \includegraphics[width=\linewidth]{PCbasis.ps} \caption{First four PC basis vectors, $\phi_i$. Absolute values are used to show the dynamic range on a logarithmic scale. Only the first three basis vectors are actually used in the emulation; any others contribute on a scale many orders of magnitude smaller.} \label{fig:PC} \end{figure} \end{center} A Gaussian process is then used to interpolate between the model results; this means that the $c-M$ relation for a new cosmology is actually a function drawn from a unit normal distribution. The covariance matrix describes the `distance' between the new model and the set of known models as given by the covariance function. The full covariance matrix, $\Sigma$, is composed of one $\Sigma_l$ for each PC, arranged along the diagonal elements such that: $\Sigma = diag(\Sigma_{1}...\Sigma_{n})$ for $n$ PCs. Each element of $\Sigma_l$ is given by: \begin{equation} \Sigma_{l;ij} = \lambda_l\prod_{k=1}^{5} \rho_{kl}^{4(\theta_{ik}-\theta_{jk})^2}, \end{equation} where $\theta_l$ represents the cosmological parameters and the $i$ and $j$ indices run over the number of models spanning the design space (in this case $i,j=1-37$), the $l$ index runs over the number of PCs and the $k$ index runs over the number of cosmological parameters. The hyperparameters, $\lambda_l, \rho_{kl}, \lambda_p$, are set by exploring the likelihood surface, which is done with a Markov chain Monte Carlo analysis, but any other algorithm that locates the maximum likelihood of a multidimensional surface could also be used. The complete expression for the posterior can be found in Equation B17 of~\cite{coyote2}. This conditions the Gaussian process to the design of the 37 models and ensures that the hyperparameters correctly capture the complexity of the surface, because they control the fit of the interpolating functions to the data. After conditioning the GP for the best-fitting hyperparameters, the emulator is ready to predict the $c-M$ relation for a different cosmology. The prediction involves re-calculating the covariance matrix between the new parameters and the design and this locates the new parameters within the design space. This process is quite fast, and can be repeated each time a new cosmology is needed with little computational cost. The results at intermediate redshifts ($0 < z < 1$) are produced with a simple linear interpolation. This remains fairly accurate because the change in the $c-M$ relation with redshift is largely a simple shift in amplitude. \subsection{Testing the Emulator}\label{subsec:testing} The accuracy of the emulator is determined using two methods: 1) we compare the performance of the emulator against a model not included in the original design and 2) we remove one of the models from the design and rebuild the emulator based on the remaining 36 models in what is known as a holdout test. In this section, we perform both of these tests to demonstrate the accuracy of the $c-M$ emulator. We withheld one model (M000) with a $\Lambda$CDM concordance cosmology from the set of 37 models when building the $c-M$ emulator. Figure~\ref{fig:M000} shows the comparison between the emulator prediction for this cosmology against the direct simulation results from three different box sizes at $z=0$ and $z=1$. The hashed region covers the 1-sigma boundary around the mean. The emulator predictions are consistent with the $N$-body $c-M$ relations well within the errors on the measurements. In comparison with the smoothed fit for model M000, derived from the same power law fitting procedure used on the set of 37 cosmologies, we find that at $z=0$, the emulator is essentially perfect at the high mass end and accurate to at least 3.25\% for low mass halos. For $z=1$ the error is somewhat worse, mainly due to the limited halo statistics for building the emulator, especially for the low-$\sigma_8$ models. At low masses the predictions are accurate at the 2\% level and degrade to 9\% inaccuracy at the highest masses considered. All of these values are well within what may be considered to be the nominal uncertainty in determining concentrations from simulations~\citep{bhattacharya11_b}. For most of the range of halo mass considered, the accuracy of the emulator outperforms any other prediction scheme available, especially considering the large model space covered here. \begin{figure}[t] \includegraphics[width=1\linewidth]{cMemu_z0.ps} \caption{Predictions from the $c-M$ emulator at $z=0$ (blue, solid) and $z=1$ (red, solid) for measurements from $N$-body simulations for the M000 cosmology. The dashed lines show the best-fit power law describing the $N$-body results and the hashed region shows the expected variation in the $c-M$ relation from the mean (solid line). Note that this set of simulations was not used to build the emulator.} \label{fig:M000} \end{figure} In Fig.~\ref{fig:holdouts}, we show estimates of the emulator error by performing a holdout test. In such a test one model is kept aside and a new emulator is built, based on the remaining 36 models. The new emulator is used to predict the $c-M$ relation for the held-out model. Since the numerical result (`truth') is known for that model, we can measure the emulator prediction error. One shortcoming of this method -- in particular if only a very small number of simulations is available as is the case here -- is that by removing one model, the quality of the emulator is degraded. Therefore, the error estimate for the emulator obtained this way can be considered to be a conservative upper bound. We have chosen to exclude only models M004, M008, M013, M016, M020 and M026, because these are located relatively close to the center of the design. Removing a model that defines one of the edges of the design would greatly reduce the performance of the emulator, since the GP would be extrapolating for a missing model that is now outside of the design range. The comparison is made with respect to the smoothed $N$-body result that was used to construct the full emulator, not the raw concentration measurements from the simulation. At most, the emulator deviates by 3.3\% from the simulation results at $z=0$ and this rises to 15\% at $z=1$. This is because the error on the raw measurements increases with redshift as the sample size of halos decreases, particularly for low-$\sigma_8$ models. \begin{figure}[t] \includegraphics[width=\linewidth]{holdout_z0.ps} \includegraphics[width=\linewidth]{holdout_z1.ps} \caption{Holdout tests for the $c-M$ emulator at $z=0$ (top) and $z=1$ (bottom). In both plots, the labelled model has been removed from the design and an emulator is rebuilt on the reduced design. We then take the ratio of the smoothed $N$-body result and the prediction from the new emulator to check the accuracy of the full emulator made with the original design.} \label{fig:holdouts} \end{figure} \subsection{Comparison with other $c-M$ Predictions} We now compare the results obtained from the emulator with those from the models presented in~\cite{bhattacharya11_b}, ~\cite{bullock},~\cite{duffy08} and~\cite{prada12}. The~\cite{bullock} model was intended to correct the redshift dependence of the original NFW model, which was claimed to overpredict the concentration of high redshift ($z > 1$) halos. We perform our comparisions against the most recent version of the model that incorporates corrections from \cite{maccio08}\footnote{available from {\tt physics.uci.edu/$\sim$bullock/CVIR}}. The~\cite{bullock} model contains two free parameters $K = 3.85$ and $F =0.01$. Newer values of $K$ and $F$ were obtained in \cite{maccio08} by fitting this model to $N$-body simulations using cosmological parameters corresponding to the first, third and fifth WMAP data releases. Figure~\ref{fig:bullock} shows the ratio of the~\cite{bullock} model to our emulator at $z=0$ for two cosmologies, M000 and WMAP7~\citep{wmap7}; note that the~\cite{bullock} model has only been tested with $\Lambda$CDM and SCDM cosmologies. These two models are certainly consistent at low halo masses, within the expected error of the emulator, but a substantial discrepancy occurs at cluster-sized halos, even with the updated version of~\cite{maccio08}. This occurs because the model contains free parameters that need to be tuned to a particular cosmology with $N$-body simulations. However, the~\cite{bullock} model is able to reach much lower halo masses, $M < 10^{10}$ M$_\odot$, than our emulator because it is calibrated to higher mass resolution $N$-body simulations. At $z=1$, the public code used for the Bullock/Macci\'o model fails to compute the concentration across the full range of halo masses because of difficulties at low $\sigma_8$. We therefore show only results for a limited mass range. The discrepancy here is much larger than for $z=0$, with a concentration underestimation of greater than 20\%. More recently,~\cite{duffy08} proposed a new $c-M$ relation with a power-law relationship between the halo mass and the concentration, as extracted from a series of high resolution, small to medium volume $N$-body simulations with a WMAP5 cosmology. Results from the $c-M$ emulator are consistent with their predictions to within $\sim$ 10\% at $z=0$, as are the results in~\cite{bhattacharya11_b}. There is a slight deviation at cluster sized halos; our $c-M$ emulator is based on larger volume simulations, and is therefore able to provide a more complete sample of massive halos and reduced shot noise at the high mass end. The agreement between the emulator and the~\cite{duffy08} $c-M$ relation improves to $\sim$ 5-6\% at $z=1$ for the WMAP5 cosmology. \begin{figure}[t] \includegraphics[width=\linewidth]{bullock.ps} \includegraphics[width=\linewidth]{bullock_z1.ps} \caption{Comparison of the emulator -- taken as the reference -- with other models for the $c-M$ relation at $z=0$ (upper panel) and $z=1$ (lower panel). The \cite{bullock}/\cite{maccio08} $c-M$ relation is shown in blue for both M000 (solid) and WMAP7 (dashed) cosmologies. The \cite{bhattacharya11_b} fit (note that this fit was derived only for M000, not for general cosmological models) is in green, the~\cite{prada12} $c-M$ relation in cyan, and the~\cite{duffy08} $c-M$ relation in red for M000 (solid) and WMAP5 (dashed) cosmologies. In addition, the pink line shows the ratio of the power-law fit for M000 to the emulator prediction (As shown in Fig.~\ref{fig:M000} the agreement is very good, with less than 3\% deviation over most of the mass range). The lower panel shows the results for $z=1$. The publicly available code for the~\cite{bullock}/\cite{maccio08} fit does not work seamlessly over the full mass range so we do not show results for the very high mass end here. See the text for further discussion of this set of results.} \label{fig:bullock} \end{figure} \begin{center} \begin{figure*} \includegraphics[width=1.75in,angle=270]{sensitivities_w_error.ps} \caption{Sensitivity of the $c-M$ relation to the five cosmological parameters varied in the emulator design, $\Omega_mh^2, \Omega_bh^2,-w, n_s$ and $\sigma_8$, at $z=0$. We vary each cosmological parameter individually for each panel and have binned the range into five intervals, which are coloured from light to dark as the value of the parameter increases. From each of these, we subtract the $c-M$ relation for the model corresponding to the midpoint of the design space, c(M)$_0$. To guide the eye, we also plot the median c(M) - c(M)$_0$ of each bin in parameter space.} \label{fig:sensitivity} \end{figure*} \end{center} In Figure~\ref{fig:bullock}, we also show the ratio between our emulator and the $c-M$ relation as determined by the model discussed in~\cite{prada12}, which is itself based on a number of $N$-body simulations. There is a $\sim$ 20\% discrepancy at $z=0$ ($\sim$ 40\% at $z=1$) for lower halo masses. This increases dramatically for cluster sized haloes at both redshifts, since unlike~\cite{prada12}, we do not observe an upturn in the $c-M$ relation, where their concentration increases with halo mass. One should note that the methods for measuring the halo concentration are different in our two cases -- we use a finite-range profile-fitting method as discussed in~\cite{bhattacharya11_b}, whereas~\cite{prada12} use a two-point ratio method. Exectations for discrepancies between profile fitting and their particular ratio method are further discussed in the appendix of~\cite{bhattacharya11_b}. We also note that there is a good agreement to within $\sim$5\% between our emulator and the $c-M$ relations measured by~\cite{neto07} from the Millennium simulation~\cite{springel05}. The cosmology of the Millennium simulation does not quite fall within the range of our emulator ($\omega_bh^2 = 0.024$ and $h = 0.73$), and to facilitate this comparsion, we have adoped values as close to these as possible that still lie within our parameter space ($\omega_bh^2 = 0.0235$ and $h = 0.719$). Lastly, we compare the emulator prediction with the fitting function derived in \cite{bhattacharya11_b} for the M000 cosmology. Before doing so, we provide some necessary background. First, the redshift dependence in the fitting form in \cite{bhattacharya11_b} is handled differently than in the current paper. In \cite{bhattacharya11_b}, the aim was to find a global power-law fit that encompasses all redshifts considered (between $z=0$ and $z=2$) at once. Therefore, the fit for each redshift is not expected to be perfect. In the current paper we follow a different path: since we do not provide a single formula for the $c-M$ relation but rather a simple numerical code, we can generate the best-fit power-law model for each redshift separately and then simply interpolate between the redshifts. This produces a more accurate answer at each redshift at the minimal cost of running a fast code for every $c-M$ prediction instead of using one fitting formula. Second, for the high-mass range, \cite{bhattacharya11_b} used higher force resolution simulations. As shown in the Appendix of \cite{bhattacharya11_b}, the concentrations from the Coyote runs are slightly lower at high masses (at the 5\% level) compared to higher-resolution simulations. Since it is not clear if this effect is independent of cosmology (most likely for lower $\sigma_8$ simulations the effect will be smaller) we decided to not attempt to correct the concentration measures in this paper for the Coyote runs. Therefore, the uncertainty for the high mass concentrations from the emulator predictions will be slightly higher and one expects the predictions to be biased slightly low. Considering the overall scatter and uncertainty in the $c-M$ relation, this small effect is unlikely to be significant. One should note, however, that due to this suppression, the ratio of $\sigma_c(M)$ to the mean concentration, which we quote at the nominal value of 1/3, could be slightly smaller at higher masses. Keeping these caveats in mind, we now turn to the comparison of the fit by \cite{bhattacharya11_b} and the emulator result, in Fig.~\ref{fig:bullock}. For $z=0$, both agree at the 2\% (low halo mass) to 6\% (high halo mass) level, the \cite{bhattacharya11_b} fit being slightly higher as expected. For $z=1$, the discrepancy ranges from 4\% to 17\%. Here, overall the emulator estimate compared to the best-fit power-law to the simulation result is slightly low, while the \cite{bhattacharya11_b} fit slightly overestimates the simulation results. In other words, the actual simulation result lies in between the emulator prediction and the fit. Overall, the agreement between the \cite{bhattacharya11_b} fit and the emulator is much better at $z=1$ than the agreement between the emulator and the Bullock/Macci\'o fit. \subsection{Cosmology Dependence of the $c-M$ Relation} Finally, we explore the sensitivity of the $c-M$ relation to variations in cosmology. Since we now have a means of quickly and smoothly interpolating from one cosmology to another, we can simply vary the parameters that we incorporated into our design space in a regular grid. We divide each parameter range into five evenly spaced regions and vary only a single parameter at a time by keeping the other four parameters fixed at the midpoint of the parameter range. Our comparisons are always made with respect to this model at the midpoint, which has the parameters: $\Omega_mh^2 = 0.1375$, $\Omega_bh^2 = 0.0225$, $n_s = 0.95$, $w = -1$ and $\sigma_8 = 0.758$, and is subtracted from each $c-M$ relation. Figure~\ref{fig:sensitivity} shows the results of this exercise at $z=0$, with the values of the cosmological parameters increasing as the shading increases from light to dark. The entire region is coloured to show the variation in the $c-M$ relation across each parameter bin. The range of concentration variation changes as a function of mass and the cosmological parameter being varied; the largest variation is of order unity. Unsurprisingly, the $c-M$ relation increases with $\Omega_mh^2$ and $\sigma_8$ and decreases with $w$, which slows the rate of structure formation. We see little variation with $\Omega_bh^2$ because the range of parameters allowed by the CMB constraints is already quite tight. Also, cluster-sized halos appear to be relatively less sensitive to these changes in cosmology. Figure~\ref{fig:sensitivity} also shows a clear degeneracy between $\Omega_mh^2$, $\sigma_8$, and $n_s$ in the $c-M$ relation. \section{Conclusion and Outlook} \label{sec:conc} In this paper, we have presented a new prediction scheme -- in the form of an emulator -- for the $c-M$ relation for dark matter-dominated halos at the bright galaxy to cluster mass scales, covering a range of $2\cdot 10^{12}$M$_\odot<M<10^{15}$M$_\odot$ and a redshift range of $z=0$ to $z=1$. The emulator provides results for a large class of $w$CDM cosmologies and is accurate at the $\sim5\%$ level (better for lower redshifts, slightly worse for higher redshifts). The emulator enables consistent predictions to be made when testing for deviations from $\Lambda$CDM using clusters. This is particularly important for cluster cosmology, since the behaviour of the $c-M$ relation can vary by as much as 30\% just by varying the equation of state across the range $-1.3 < w<-0.7$. By correctly including the cosmology dependence in the $c-M$ relation, the emulator improves on analytic modelling of halo profiles, such as the 1-halo term used in the halo power spectrum. The performance of the emulator compares favourably with the other models for the $c-M$ relation in the literature and outperforms the Bullock/Macci\'o model across the redshift and mass range considered. Aside from predicting the mean $c-M$ relation, the interesting and useful fact that across all 37 cosmologies considered, 1) the scatter in halo concentrations in individual mass bins is Gaussian, and 2) the corresponding standard deviation is given by roughly a third of the mean concentration value, means that the $c-M$ emulator also includes within it the information regarding the concentration distribution at a given value of mass. The work in this paper is an example of how the cosmic calibration framework provides a means of estimating highly nonlinear quantities involving evolved structures from a limited number of computationally expensive $N$-body simulations. In the future, we will extend the number and range of cosmological parameters to include more exotic phenomena, such as evolving dark energy, as a complement to upcoming dark energy experiments. \begin{acknowledgments} The work at Argonne National Laboratory was supported under U.S. Department of Energy contract DE-AC02-06CH11357. Part of this research was supported by the DOE under contract W-7405-ENG-36. We are indebted to Charlie Nakhleh for providing an example code for building cosmic emulators and Earl Lawrence and Dave Higdon for many useful and entertaining discussions on the topic. We thank Volker Springel for making {\sc GADGET-2} publicly available. We are grateful for computing time granted to us as part of the Los Alamos Open Supercomputing Initiative. This research used resources of the Argonne Leadership Computing Facility at Argonne National Laboratory and the National Energy Research Scientific Computing Center, which are supported by the Office of Science of the U.S. Department of Energy under contract DE-AC02-06CH11357 and DE-AC02-05CH11231 respectively. \end{acknowledgments} \newpage
\section{Introduction}\label{s-intro} Subdwarf B (sdB) stars are considered to be core helium burning stars with a very thin (M$_{\rm{H}}$ $<$ 0.02 $M_{\odot}$) hydrogen envelope, and a mass close to the core helium flash mass of 0.47 $M_{\odot}$. Their existence provides stellar evolutionary theory with a challenge: how can a red giant lose its entire hydrogen envelope before starting core helium fusion? \citet{Mengel75} was the first to propose the hypothesis of binary evolution to form sdB stars. Currently there are five evolutionary channels that are thought to produce sdB stars, all of which are binary evolution channels where interaction physics play a major role \citep{Heber09}. The first two are common envelope (CE) ejection channels. In the first of these, a CE is formed by unstable mass transfer and expelled due to energy transfer from the shrinking orbit to the envelope. In the second CE channel, the envelope is not ejected in the initial dynamical phase, but at a later stage, when the companion penetrates the initial radiative layer of the giant. These channels produce short-period sdB binaries with $P_{\rm{orb}}$ = 0.1 -- 10 d. Furthermore, there are two stable Roche-lobe overflow (RLOF) channels. In the first, distinct mass transfer starts near the tip of the first giant branch, and in the second, mass transfer occurs when the binary is passing through the Hertzsprung gap. Subdwarf B binaries formed in this way are predicted to have longer periods ranging from 10 to over 500 days. All of the above scenarios result in a very small mass range for the sdB component, peaking at $M_{\rm{sdB}}$ = 0.47 $\pm$ 0.05 $M_{\odot}$. The last channel that can produce an sdB star is the double white dwarf merger channel, where a pair of close helium-core white dwarfs loses orbital energy through gravitational waves, resulting in a merger. This channel forms a single subdwarf star with a higher mass, up to 0.65 $M_{\odot}$ \citep{Webbink84}. The binary interaction processes that form these sdB binaries are very complex, and many aspects are still not completely understood. Formally, they are described using several parameters such as efficiency of envelope ejection, accretion efficiency, physical interaction during common envelope phase, etc. Most of these are currently poorly constrained by observations. The most common tool to constrain evolution paths is by performing a binary population synthesis (BPS) and comparing the resulting population to an observed population for different values of these parameters. \citet{Han02, Han03}, \citet{Nelemans10} and \citet{Clausen12} have performed such BPS studies. The main results are that subdwarf progenitors with a low-mass companion will have unstable mass transfer on the RGB, during which their orbit will shrink, resulting in short-period sdB binaries. Subdwarf progenitors with a more massive companion will have expanded orbits, resulting in long-period sdB binaries. Several observational studies \citep{Maxted01,Morales03, Copperwheat11} have focused on short-period binaries, and found that about half of all sdB stars reside in short-period (P$_{\rm{orb}}$ $<$ 10 d) binaries. Longer period sdB binaries have been observed \citep[e.g.][]{Green01}, but no definite orbits have been established. \citet{Clausen12} concluded that, although the currently known population of sdB binaries with a white dwarf (WD) or M dwarf companion can constrain some parameters used in the binary evolution codes, many parameter remain open. Furthermore they concluded that sdB binaries with a main-sequence (MS) component will be able to provide a better understanding of the limiting mass ratio for dynamically stable mass transfer during the RLOF phase, the way that mass is lost to space, and the transfer of orbital energy to the envelope during a common envelope phase. The period distribution of sdB + MS binaries can support the existence of stable mass transfer on the RGB. If long-period sdB + MS binaries are found, they are thought to be formed during a stable mass transfer phase in the RLOF evolutionary channel. If no long-period sdB + MS binaries can be found, this would indicate that mass transfer on the RGB is unstable. The goal of this paper is to describe the methods we used to determine radial velocities of both the main-sequence and subdwarf B component in sdB + MS systems (Sect.\,\ref{s-spect}), and the use of spectral energy distributions (SEDs) to determine the spectral type of both components (Sect.\,\ref{s-sed}). Furthermore, if the mass ratio is known from spectroscopy, the SEDs from high-precision broad-band photometry alone can be used to determine their surface gravities with an accuracy up to $\Delta\log{g}$ = 0.05 dex (cgs) and the temperatures with an accuracy as high as 5\%. The resulting surface gravities are independently confirmed by the surface gravities derived from the gravitational redshift (Sect.\,\ref{s-gr}). After which all obtained parameters are summarized (Sect.\,\ref{s-absolutedimensions}). The equivalent widths of iron lines in the spectra can be used to independently confirm the atmospheric parameters of the cool companion (Sect.\,\ref{s-ironlines}). These methods were applied to PG\,1104+243. The cool companion of PG\,1104+243 is presumed to be on the main-sequence, and we will refer to it as the MS component in this paper. PG\,1104+243 is part of a long-term spectroscopic observing program, and preliminary results of this and seven more systems in this program were presented by \citet{Oestensen11, Oestensen12}. \section{Spectroscopy}\label{s-spect} The high-resolution spectroscopy of PG\,1104+243 was obtained with the HERMES spectrograph (High Efficiency and Resolution Mercator Echelle Spectrograph, R = 85000, 55 orders, 3770-9000 \AA, \citealt{Raskin11}) attached to the 1.2--m Mercator telescope at the Roque de los Muchachos Observatory, La Palma. HERMES is isolated in an over pressurized temperature-controlled enclosure for optimal wavelength stability, and connected to the Mercator telescope with an optical fiber. In total 38 spectra of PG\,1104+243 were obtained between January 2010 and February 2012. The date and exposure time of each spectrum is shown in Table\,\ref{tb-observations}. HERMES was used in high-resolution mode, and Th-Ar-Ne exposures were made at the beginning and end of the night, with the exposure taken closest in time used to calibrate the wavelength scale. The exposure time of the observations was adapted to reach a signal-to-noise ratio (S/N) of 25 in the $V$--band. The HERMES pipeline v3.0 was used for the basic reduction of the spectra, and includes barycentric correction. Part of a sample spectrum taken with HERMES is shown in Fig.\,\ref{fig-spectrum}. Furthermore, five intermediate-resolution spectra (R = 4100) were obtained with the Blue Spectrograph at the Multi Mirror Telescope (MMT) in 1996-1997. The Blue Spectrograph was used with the 832/mm grating in 2nd order, covering the wavelength region 4000-4950 \AA. The data were reduced using standard IRAF tasks. An overview of the spectra is given in Table\,\ref{tb-rv_green}. \begin{figure*}[!t] \centering \includegraphics{f_spectrum.eps} \caption{A sample normalized spectrum of PG\,1104+243 (black), showing the 5875 \AA\ He\,I line of the sdB component, and several lines as e.g. the Na doublet at 5890-5896 \AA\ of the MS component. Furthermore, a synthetic spectrum of the MS component (green) and sdB component (red) are shown.}\label{fig-spectrum} \end{figure*} \begin{table} \caption{The observing date (mid-time of exposure), exposure time and the initials of the observer of the PG\,1104+243 spectra observed with HERMES.} \label{tb-observations} \centering \begin{tabular}{llllll} \hline\hline BJD & Exp. & Obs. & BJD & Exp. & Obs. \\ --2450000 & s & & --2450000 & s & \\\hline 5204.70741 & 1400 & K.S. & 5652.55549 & 1300 & P.N. \\ 5204.72527 & 1400 & K.S. & 5655.60536 & 1400 & P.N. \\ 5217.67376 & 1500 & C.W. & 5658.47913 & 1400 & P.N. \\ 5217.69173 & 1500 & C.W. & 5666.43279 & 1200 & S.B. \\ 5234.60505 & 900 & K.E. & 5666.45153 & 1200 & S.B. \\ 5234.61608 & 900 & K.E. & 5672.53688 & 1500 & S.B. \\ 5234.62709 & 900 & K.E. & 5685.44967 & 1200 & R.L. \\ 5264.54423 & 2700 & T.M. & 5718.44565 & 1600 & A.T. \\ 5340.42820 & 1200 & R.L. & 5914.76439 & 1500 & P.L. \\ 5351.37760 & 1500 & S.B. & 5937.71438 & 1800 & C.W. \\ 5553.70058 & 1200 & K.E. & 5943.63769 & 1800 & N.G. \\ 5569.69768 & 1200 & S.B. & 5953.76375 & 1200 & J.M. \\ 5579.56553 & 1500 & P.D. & 5957.72053 & 1200 & J.M. \\ 5579.58347 & 1500 & P.D. & 5959.59752 & 1200 & J.M. \\ 5589.75558 & 1200 & P.L & 5963.64527 & 3600 & J.V. \\ 5611.64081 & 1200 & P.C. & 5964.53128 & 3500 & J.V. \\ 5622.61267 & 1200 & N.C. & 5964.65639 & 2400 & J.V. \\ 5639.58778 & 1500 & N.G. & 5966.69796 & 2400 & J.V. \\ 5650.50737 & 1500 & B.A. & 5968.61334 & 1800 & J.V. \\ \hline \end{tabular} \end{table} To check the wavelength stability of HERMES, 38 different radial velocity standard stars of the IAU were observed over a period of 1481 days, one standard star a night, coinciding with the observing period of PG1104+243. These spectra are cross-correlated with a line mask corresponding to the spectral type of each star to determine the radial velocity. For this cross-correlation (CC), the {\tt hermesVR} method of the HERMES pipeline, which is designed to determine the radial velocity of single stars, is used. This method handles each line separately, starting from a given line-mask. The method uses the extracted spectra, after the cosmic clipping was performed but prior to normalization. It performs a cross-correlation for each line in the mask, and sums the CC functions in each order. The radial velocity is derived by fitting a Gaussian function to the CC function \citep{Raskin11}. To derive the final radial velocity only orders 55$\rightarrow$74 (5966 -- 8920 \AA) are used, as they give the best compromise between maximum S/N for G-K type stars and, after masking the telluric bands, absence of telluric influence \citep{Raskin11}. The resulting standard deviation of these radial velocity measurements is 80 m/s with a non significant shift to the IAU radial velocity standard scale. \subsection{Radial Velocities}\label{s-radialvelocities} The determination of the radial velocities of PG\,1104+243 from the HERMES spectra was performed in several steps. First, there were several spectra taken in a short period of time. As the orbital period of PG\,1104+243 was determined at 752 $\pm$ 14 d by \citet{Oestensen12}, spectra that were taken on the same night, or with only a couple of nights in between are summed to increase the S/N. We experimented with several different intervals, and determined that when spectra taken within a five-day interval (corresponding to 0.7\% of the orbital period, or a maximum radial velocity shift of 0.05 km/s) were merged, there was no significant smearing or broadening of the spectral lines. After this merging, 25 spectra with a S/N ranging from 25 to 50 remained. The MS component has many spectral lines in every Echelle order, most of which are not disturbed by the He lines of the sdB component. Because of this, the CC-method of the HERMES pipeline ({\tt hermesVR}) can be used to determine the radial velocities of the MS component without any loss of accuracy. For the MS component of PG\,1104+243, a G2-type mask was used. The final errors on the radial velocities take into account the formal errors on the Gaussian fit to the normalized cross-correlation function, the error due to the stability of the wavelength calibration and the error arising due to the used mask. The resulting velocities and their errors are shown in Table\,\ref{tb-rv}. The determination of the radial velocities of the sdB component is more complicated. Except for the Balmer lines, there are only a few broad He lines visible in the spectrum. After comparison of a synthetic spectrum of a G2-type star with the spectra, it was found that only the 5875.61 \AA\ He\,I blend is not contaminated by lines of the MS component. For this single blend, we had to deploy another method to obtain an accurate radial velocity. As only one line is used, anomalies in the spectra can cause significant errors in the radial velocity determination. To prevent this the region around the He\,I line is cleaned from any remaining cosmic rays by hand. The normalization process is designed to run automatically, making it possible to quickly process multiple lines is many spectra, and consists of two steps. In the first, a spline function is fitted through a 100 \AA\ region centered at the He\,I line to remove the response curve of HERMES, and the main features of the continuum. In the second step, a small 10 \AA\ region around the He\,I line is further normalized by fitting and re-fitting a low degree polynomial through the spectrum. After each iteration, the flux points lower than the polynomial are discarded. This is repeated until the polynomial fits the middle of the continuum. An example of a spectrum normalized using this method is shown in Fig.\,\ref{fig-spectrum}. These cleaned and normalized spectra can then be cross-correlated to a synthetic sdB spectrum. For this purpose a high-resolution LTE spectrum of $T_{\rm{eff}}$ = 35000 K and $\log{g}$ = 5.50 from the grids of \citet{Heber00} was used. The resolution of this synthetic spectrum is matched to the resolution of the HERMES spectra. Synthetic spectra with different temperatures and surface gravities were tested, to check the temperature and $\log{g}$ dependence of the radial velocities. Changing either of those parameters causes a systemic shift of maximum $0.2\ km/s$ in the resulting radial velocities. The cross-correlation is performed in wavelength space by starting at the expected radial velocity, calculated from the radial velocities of the MS component. To determine the radial velocity, a Gaussian is fitted to the cross-correlation function. The error on the radial velocities is obtained from a Monte-Carlo simulation. This is done by adding Gaussian noise to the spectrum and repeating the cross-correlation with the synthetic spectrum. The level of the Gaussian noise is determined from the noise level in the continuum near the He\,I line. The final error is determined by the standard deviation from the radial velocity results of 1000 simulations, the wavelength stability of HERMES and the dependence on the template. The resulting radial velocities and their errors are shown in Table\,\ref{tb-rv}. To derive the radial velocities for the MS component from the MMT spectra, the spectra were cross-correlated against two high S/N G0V spectra (HD13974 and HD39587) using the IRAF task {\tt FXCOR}, where the cross-correlation was done only in the wavelength ranges containing absorption lines from the MS companion, ie avoiding the Balmer lines, He\,$\lambda$ 4026, He\,$\lambda$ 4387, He\,$\lambda$ 4471, He\,$\lambda$ 4686, He\,$\lambda$ 4712, and He\,$\lambda$ 4921. The velocities derived from the two G0V stars were averaged to get the velocity of the MS companion for each of the five PG\,1104+243 spectra. The proper scaling factor for the velocity errors output by {\tt FXCOR} was determined using the standard deviation of the velocities obtained from each template relative to the average velocity for each epoch. The resulting radial velocities and errors and are given in Table\,\ref{tb-rv_green}. \begin{table} \caption{The radial velocities of both components of PG\,1104+243.} \label{tb-rv} \centering \begin{tabular}{lrrrr} \hline\hline \noalign{\smallskip} & \multicolumn{2}{r}{MS component} & \multicolumn{2}{r}{sdB component} \\ BJD & RV & Error & RV & Error \\ -2450000 & km s$^{-1}$ & km s$^{-1}$ & km s$^{-1}$ & km s$^{-1}$ \\\hline \noalign{\smallskip} 5204.7163 & -12.014 & 0.146 & -19.25 & 1.18 \\ 5217.6827 & -12.036 & 0.150 & -20.32 & 0.65 \\ 5234.6160 & -11.323 & 0.147 & -19.66 & 0.23 \\ 5264.5442 & -11.345 & 0.147 & -20.87 & 0.48 \\ 5340.4282 & -11.916 & 0.146 & -19.37 & 0.57 \\ 5351.3776 & -12.138 & 0.149 & -19.19 & 0.56 \\ 5553.7005 & -18.106 & 0.147 & -9.31 & 0.39 \\ 5569.6976 & -18.679 & 0.150 & -8.91 & 0.26 \\ 5579.5745 & -19.183 & 0.149 & -8.54 & 0.36 \\ 5589.7555 & -19.063 & 0.150 & -8.37 & 0.29 \\ 5611.6408 & -19.607 & 0.149 & -7.28 & 0.22 \\ 5622.6126 & -19.982 & 0.146 & -6.74 & 0.54 \\ 5639.5877 & -19.622 & 0.146 & -6.76 & 0.52 \\ 5651.5314 & -20.112 & 0.144 & -6.96 & 0.35 \\ 5657.0422 & -19.873 & 0.146 & -6.25 & 1.02 \\ 5666.4421 & -20.327 & 0.144 & -6.89 & 1.32 \\ 5672.5368 & -20.259 & 0.147 & -7.25 & 0.23 \\ 5685.4496 & -20.119 & 0.146 & -6.81 & 0.48 \\ 5718.4456 & -19.705 & 0.147 & -6.95 & 0.34 \\ 5914.7644 & -13.405 & 0.147 & -16.88 & 0.38 \\ 5937.7143 & -12.741 & 0.144 & -18.19 & 0.22 \\ 5943.6376 & -12.528 & 0.146 & -18.58 & 0.48 \\ 5955.7421 & -12.361 & 0.147 & -18.94 & 0.82 \\ 5962.5913 & -12.214 & 0.144 & -18.84 & 0.71 \\ 5966.6559 & -11.994 & 0.147 & -19.21 & 0.56 \\ \hline \end{tabular} \end{table} \begin{table} \caption{The observing date (mid-time of exposure), exposure time and radial velocities of the MS component of PG\,1104+243, determined from the spectra observed with the Blue Spectrograph at the MMT.} \label{tb-rv_green} \centering \begin{tabular}{lrrrr} \hline\hline \noalign{\smallskip} BJD & Exp. & RV & Error \\ -2450000 & s & km s$^{-1}$ & km s$^{-1}$\\\hline \noalign{\smallskip} 435.0168 & 300 & -20.21 & 0.60 \\ 436.0515 & 300 & -19.21 & 0.63 \\ 476.9221 & 400 & -18.24 & 0.66 \\ 510.8811 & 120 & -18.94 & 0.61 \\ 836.0087 & 200 & -11.94 & 0.58 \\ \hline \end{tabular} \end{table} \subsection{Orbital Parameters}\label{s-orbitalparamters} The orbital parameters were calculated by fitting a Keplerian orbit through the radial-velocity measurements, while adjusting the period ($P$), time of periastron ($T_0$), eccentricity ($e$), angle of periastron ($\omega$), amplitude ($K$) and systemic velocity ($\gamma$), using the orbital period determined by \citet{Oestensen12} as a first guess for the period. The MS and sdB component were treated separately in this procedure. The radial velocities of both components were weighted according to their errors as $w = 1/\sigma$. Fitting the orbit resulted in a very low eccentricity ($e < 0.002$). To check if the orbit of the sdB+MS binary is circularized, the \citet{Lucy71} test for circularized orbits was applied: \begin{equation} P = \left( \frac{\sum{ (o-c)^2_{\rm{ecc}}} }{\sum{ (o-c)^2_{\rm{circ}}} } \right)^{(n-m)/2}, \end{equation} where ecc indicates the residuals of an eccentric fit, and circ the residuals of a circular fit, $n$ the total number of observations, and $m = 6$, the number of free parameters in an eccentric fit. If $P < 0.05$, the orbit is significantly eccentric, if $P \geq 0.05$, the orbit is effectively circular. In the case of PG\,1104+243 we find P = 0.15, indicating a circular orbit. This is expected as the orbit is supposed to be circularized by tidal interactions during the RGB evolution \citep{Zahn77}. In the further determination of the orbital parameters, a circular orbit is assumed. To derive the final orbital parameters, both components were first treated separately to measure the difference in systemic velocity (see Sect.\,\ref{s-gr}). Then that difference was subtracted from the radial velocities of the sdB component, and both components are solved together to obtain their spectroscopic parameters. The uncertainties on the orbital parameters were determined by using Monte Carlo-simulations. The radial velocities were perturbed based on their errors, and the errors on the parameters were determined by their standard deviation after 5000 iterations. The residuals of the orbital fit to the MS component are larger than the errors on the radial velocities. However, this variability is most likely intrinsic, caused by stellar spots or low amplitude pulsations, and does not affect the derived orbital parameters. The spectroscopic parameters of PG\,1104+243 are shown in Table\,\ref{tb-specparam}. The radial-velocity curves and the best fit solution are plotted in Fig \ref{fig-rvcurves}. \begin{table} \centering \caption{Spectroscopic orbital solution for both the main-sequence (MS) and subdwarf B (sdB) component of PG\,1104+243.} \label{tb-specparam} \begin{tabular}{lrr} \hline\hline \noalign{\smallskip} Parameter & \multicolumn{1}{c}{MS} & \multicolumn{1}{c}{sdB} \\\hline \noalign{\smallskip} $P$ (d) & \multicolumn{2}{c}{753 $\pm$ 3} \\ $T_0$ & \multicolumn{2}{c}{2450386 $\pm$ 8} \\ $e$ & \multicolumn{2}{c}{$<$ 0.002} \\ $q$ & \multicolumn{2}{c}{0.64$\pm$0.01} \\ $\gamma$ (km s$^{-1}$) & $-$15.63 $\pm$ 0.06 & $-$13.7 $\pm$ 0.2 \\ $K$ (km s$^{-1}$) & 4.42 $\pm$ 0.08 & 6.9 $\pm$ 0.2 \\ $a$ $\sin{i}$ ($R_{\odot}$) & 66 $\pm$ 1 & 103 $\pm$ 1 \\ $M$ $\sin^3{i}$ ($M_{\odot}$) & 0.069 $\pm$ 0.002 & 0.044 $\pm$ 0.002 \\ $\sigma$ & 0.18 & 0.27 \\ \hline \end{tabular} \tablefoot{$a$ denotes the semi-major-axis of the orbit. The quoted errors are the standard deviation from the results of 5000 iterations in a Monte Carlo simulation.} \end{table} \begin{figure} \centering \includegraphics{f_rv_curves.eps} \caption{Top: spectroscopic orbital solution for PG\,1104+243 (solid line: MS, dashed line: sdB), and the observed radial velocities (blue filled circles: MS HERMES spectra, blue open circles: sdB HERMES spectra, green triangles: MS MMT spectra). The measured system velocity of both components is shown by a dotted line. Middle: residuals of the MS component. Bottom: residuals of the sdB component.} \label{fig-rvcurves} \end{figure} \section{Spectral energy distribution}\label{s-sed} To derive the spectral type of the main-sequence and subdwarf component, we fitted the photometric spectral energy distribution (SED) of PG\,1104+243 with model SEDs. With this method we can determine the effective temperature and surface gravity of both components with good accuracy. \subsection{Photometry} Photometry of PG\,1104+243 was collected using the subdwarf database\footnote{http://catserver.ing.iac.es/sddb/} \citep{Oestensen06}, which contains a compilation of data on hot subdwarf stars collected from the literature. In total, fifteen photometric observations were found in four different systems: Johnson $B$ and $V$, Cousins $R$ and $I$, Str\"{o}mgren $uvby$, and 2MASS $J$, $H$ and K$_s$. Four observations have uncertainties larger or equal than 0.1 mag and are discarded. The observations used in the SED fitting process are shown in Table\,\ref{tb-photometry}. Accurate photometry at both short and long wavelengths are used to establish the contribution of both the hot sdB component and the cooler MS component. \begin{table} \caption{Photometry of PG\,1104+243 collected from the literature.} \label{tb-photometry} \centering \begin{tabular}{lrrrr} \hline\hline \noalign{\smallskip} Band & Wavelength & Width & Magnitude & Error \\ & \AA & \AA & mag & mag \\\hline \noalign{\smallskip} Johnson $B$\tablefootmark{a} & 4450 &940 & 11.368 & 0.010 \\ Johnson $V$\tablefootmark{a} & 5500 &880 & 11.295 & 0.013 \\ Cousins $R$\tablefootmark{a} & 6500 &1380 & 11.161 & 0.010 \\ Cousins $I$\tablefootmark{a} & 7880 &1490 & 11.001 & 0.010 \\ Str\"{o}mgren $u$\tablefootmark{b} & 3460 &300 & 11.513 & 0.045 \\ Str\"{o}mgren $v$\tablefootmark{b} & 4100 &190 & 11.485 & 0.045 \\ Str\"{o}mgren $b$\tablefootmark{b} & 4670 &180 & 11.334 & 0.045 \\ Str\"{o}mgren $y$\tablefootmark{b} & 5480 &230 & 11.256 & 0.006 \\ 2MASS $J$\tablefootmark{c} & 12410 &1500 & 10.768 & 0.026 \\ 2MASS $H$\tablefootmark{c} & 16500 &2400 & 10.520 & 0.027 \\ 2MASS $K_s$\tablefootmark{c} & 21910 &2500 & 10.510 & 0.023 \\ \hline \end{tabular} \tablebib{ \tablefoottext{a}{\citet{Allard94}} \tablefoottext{b}{\citet{Wesemael92}} \tablefoottext{c}{\citet{Skrutskie06}} } \end{table} \subsection{SED fitting} In the SED of PG\,1104+243 in Fig. \ref{fig-sed}, we see both the steep rise in flux towards the shorter wavelengths of the sdB component and the bulk in flux in the red part of the SED of the MS component. To fit a synthetic SED to the observed photometry, Kurucz atmosphere models \citep{Kurucz93} are used for the MS component, and TMAP (T\"{u}bingen NLTE Model-Atmosphere Package, \citealt{Werner03}) atmosphere models for the sdB component. The Kurucz models used in the SED fit have a temperature range from 4000 to 9000 K, and a surface gravity range of $\log{g}$=3.5 dex (cgs) to 5.0 dex (cgs). The TMAP models we have used cover a temperature range from 20000 K to 50000 K, and $\log{g}$ from 4.5 dex (cgs) to 6.5 dex (cgs). They are calculated using an atmospheric mixture of 97 \% hydrogen and 3 \% helium in mass due to the expected He-depletion in the atmosphere of sdB stars. The original TMAP atmosphere models cover a wavelength range of 2500--15000 \AA. To include the 2MASS photometry as well, they are extended with a black body of corresponding temperature to 24000 \AA. The SEDs are fitted following the grid-based method described in \citet{Degroote11}, but we extended it to include constraints from binarity. In the binary scenario, there are eight free parameters to consider; the effective temperatures ($T_{\rm{eff,MS}}$ and $T_{\rm{eff,sdB}}$), surface gravities ($g_{\rm{MS}}$ and $g_{\rm{sdB}}$) and radii ($R_{\rm{MS}}$ and $R_{\rm{sdB}}$) of both components. The interstellar reddening $E(B-V)$ is naturally presumed equal for both components, and is incorporated using the reddening law of \citet{Fitzpatrick2004}. The models are first corrected for interstellar reddening, and then integrated over the photometric passbands. The distance ($d$) to the system, acts as a global scaling factor. The radii of both components are necessary as scale factors for the individual fluxes when combining the atmosphere models of both components to a single SED. Including the distance to the system, the flux of the combined SED is given by: \begin{equation} F_{tot}(\lambda) = \frac{1}{d^2} \left( R_{\rm{MS}}^2 F_{\rm{MS}}(\lambda) + R_{\rm{sdB}}^2 F_{\rm{sdB}}(\lambda) \right),\label{e-sumflux} \end{equation} where $F_{\rm{MS}}$ is the flux of the MS model, and $F_{\rm{sdB}}$ is the flux of the sdB model, in a particular passband. However, if the masses of both components are known, the radii can be derived and removed as free parameters. Assuming that the mass of the sdB component is $M_{\rm{sdB}}$ = 0.47 $M_{\odot}$ as predicted by stellar evolution models (see Sect. \ref{s-intro}), the mass of the MS component can be calculated using the mass ratio derived from the radial velocity curves (see Sect. \ref{s-orbitalparamters}): $M_{\rm{MS}} = 1/q\ M_{\rm{sdB}}$. For each model in the grid, the radii of both components are derived from their surface gravities and masses, according to: \begin{equation} R_i = \sqrt{\frac{G M_i}{g_i}},\: i=\rm{sdB, MS}\label{e-loggR}, \end{equation} in which $G$ is the gravitational constant. This relation can be used to eliminate the radius dependence in Eq. \ref{e-sumflux}: \begin{equation} F_{tot}(\lambda) = \frac{G\ M_{\rm{MS}}}{d^2\ g_{\rm{MS}}} \left( F_{\rm{MS}}(\lambda) + q \frac{g_{\rm{MS}}}{g_{\rm{sdB}}}\ F_{\rm{sdB}}(\lambda) \right). \end{equation} The distance to the system is computed by shifting the combined synthetic model flux ($F_{tot}$) to the photometric observations. As the effect of the mass of the main-sequence component ($M_{\rm{MS}}$) in this equation can be adjusted by the distance, the resulting flux is only dependent on the mass ratio of the components, and not the presumed mass of the sdB component. By using this mass ratio as a limiting factor on the radii of both components, the number of free parameters in the SED fitting process is reduced from eight to six. To select the best model, the $\chi^2$ value of each fit is calculated using the sum of the squared errors weighted by the uncertainties $(\xi_i)$ on the observations: \begin{equation} \chi^2 = \sum_i{ \frac{(O_i-C_i)^2}{\xi_i^2}}, \end{equation} with $O_i$ the observed photometry and $C_i$ the calculated model photometry. The expectation value of this distribution is $k = \rm{N}_{\rm{obs}} - \rm{N}_{\rm{free}}$, with N$_{\rm{obs}}$ the number of observations and N$_{\rm{free}}$ the number of free parameters in the fit. In our case, $k=11-6=5$. Based on this $\chi^2$ statistics the error bars on the photometry can be checked. If the obtained $\chi^2$ is much higher than the expectation value, the errors on the photometry are underestimated. If on the other hand, the obtained $\chi^2$ is much lower than the expectation value, the errors on the photometry are overestimated or we are over fitting the photometry. If the $\chi^2$-statistics are valid, the uncertainties on the fit parameters can be estimated using the cumulative density function (CDF) of the statistics. To calculate the CDF, the obtained $\chi^2$ values have to be rescaled. The $\chi^2$ of the best fitting model is shifted to the expectation value of the $\chi^2$-distribution, and all other $\chi^2$ values in the grid are scaled relative to the best $\chi^2$. The probability of a model to obtain a certain $\chi^2$ value is given by the CDF, which can be computed using: \begin{equation} F(\chi^2,k) = P\left(\frac{k}{2}, \frac{\chi^2}{2}\right) \label{e-cdf} \end{equation} Where P is the regularized $\Gamma$-function. Based on this CDF the uncertainties on the parameters can be calculated based on the distribution of the probabilities. \subsection{Results} To fit the SED of PG\,1104+243, first a grid of composite binary spectra is calculated for $\sim$2\,000\,000 points randomly distributed in the $T_{\rm{eff}}$, $\log{g}$ and E(B-V) intervals of both components. These intervals are then adjusted based on the 95\% probability intervals of the first fit, and the fitting process is repeated. To determine final confidence intervals, separate grids of $\sim$1\,000\,000 points which only vary two parameters, while keeping the other parameters fixed at their best fit values, are used. The uncertainties on the parameters are determined based on the 95\% probability intervals. The best fit has a reduced $\chi^2$ of 6.3, and with five degrees of freedom, the $\chi^2$ values in the grid were rescaled by a factor 1.26 which is equivalent to slightly increasing the photometric uncertainties. The SED fit resulted in an effective temperature of 5930 $\pm$ 160 K and 33500 $\pm$ 1200 K for the MS and sdB components respectively, while the surface gravity was determined at respectively 4.29 $\pm$ 0.05 dex and 5.81 $\pm$ 0.05 dex for the MS and sdB component. The reddening of the system was found to be 0.001$^{+0.017}_{-0.001}$ mag. This reddening can be compared to the dust map of \citet{Schlegel98}, which gives an upper bound of E(B-V) = 0.018 for the reddening in the direction of PG\,1104+243. The reddening of E(B-V) = 0.001$^{+0.017}_{-0.001}$ mag indicates that PG\,1104+243 is located in front of the interstellar molecular clouds. This is confirmed spectroscopically, as there are no sharp interstellar absorption lines (e.g. Ca {\sc ii} K-H, K {\sc i}) visible in the spectra. These parameters indicate that the cool companion is a G-type star. The results of the fit together with the probability intervals are shown in Table\,\ref{tb-sedresults}. The optimal SED fit is plotted in Fig.\,\ref{fig-sed}, while the grids with the probability for each grid point are plotted in Fig.\,\ref{fig-grid}. The probability distribution for the surface gravity and effective temperature of the MS component follows a Gaussian pattern, while the probability distribution for the same parameters of the sdB component is elongated. This latter effect is related to a moderate correlation between $\log{g}$ and $T_{\rm{eff}}$. Thus, the effect on the models caused by an increase in $T_{\rm{eff}}$, can be diminished by increasing $\log{g}$ as well. This correlation is stronger towards shorter wavelengths, so that the main-sequence component, which is mainly visible in the red part of the SED, is less affected. A similar effect is visible in the relation between $E(B-V)$ and $T_{\rm{eff}}$. \begin{figure} \centering \includegraphics{f_sed.eps} \caption{Top: The spectral energy distribution of PG\,1104+243. The measurements are given in blue, the integrated synthetic models are shown in black, where a horizontal error bar indicates the width of the passband. The best fitting model is plotted in red. Bottom: The (O-C) value for each synthetic flux point.} \label{fig-sed} \end{figure} \begin{figure*}[!t] \centering \includegraphics{f_grid.eps} \caption{The confidence intervals of the SED fit of PG\,1104+243. The upper row show distributions for the main-sequence component, the lower row for the subdwarf component. The best fit results are indicated with a red cross. The different colors show the cumulative density probability connected to the $\chi^2$ statistics given by equation \ref{e-cdf}, with the $\chi^2$ values of the grids rescaled by a factor of 1.26.}\label{fig-grid \end{figure*} As a check of the results, a SED fit of PG\,1104+243 without any assumptions on the mass of both components is performed as well. In this case, the radii of the components are randomly varied between 0.5 $R_{\odot}$ and 2.0 $R_{\odot}$ for the MS component and between 0.05 $R_{\odot}$ and 0.5 $R_{\odot}$ for the sdB component. The results of this fit are given in the lower half of Table\,\ref{tb-sedresults}. The resulting best fit parameters are very close to the results from the mass ratio-constrained fit, but apart from the uncertainty on the MS effective temperature, the uncertainties on the parameters are much larger. For both the MS and sdB surface gravity, the 95\% probability interval are larger than the range of the models. The effective temperature and surface gravity of both components of PG\,1104+243 correspond with the ionization balance of the iron lines seen in the spectrum of MS component, and the Balmer lines of the sdB component. The reason for the high accuracy of the mass ratio-constrained fit compared to a fit in which the radii are unconstrained is shown in Fig.\,\ref{fig-models_sdb}. Here five model SEDs are shown, calculated using the best fit parameters from Table\,\ref{tb-sedresults}, but with varying surface gravity, and accordingly adjusted radii, for the sdB component. The best fit model with $\log{g} = 5.81$ dex (cgs) is shown in blue, while models with a $\log{g}$ at the edge of the confidence interval ($\log{g}$ = 5.77 and 5.85 dex) are plotted in black and green. When varying $\log{g}$, the radius of the sdB component changes, while the radius of the MS component remains constant. It is this radius change that causes the large deviation in the models, and allows for an accurate determination of the surface gravity. When only the gravity is changed, and the radii of both components is kept constant, the difference in absolute magnitude in the Str\"{o}mgren-u band is of the order of $10^{-4}$ mag, while when the radius is adapted using the mass, the difference is on the order of $10^{-2}$ mag. Furthermore, because the absolute flux depends on the radius, this method has the advantage that all photometric observations are used to constrain the $\log{g}$ of each component, instead of only the observations in the blue or the red part. This analysis of PG\,1104+243 clearly shows the power of binary SED fitting when the mass ratio of the components is known. \begin{table} \caption{The results of the SED fit, together with the 95\% and 80\% probability intervals. In the upper part (fixed radius) the results when the radii of both components are determined from the $\log{g}$ and mass, are shown. In the lower part (free radius) the results with unconstrained radii are shown. For some parameters the 95\% and 80\% probability intervals could not be determined as they are larger than the range of the models.} \label{tb-sedresults} \centering \begin{tabular}{lrr@{--}lr@{--}l} \hline\hline \noalign{\smallskip} Parameter & Best fit & \multicolumn{2}{c}{95\%} & \multicolumn{2}{c}{80\%} \\\hline \noalign{\smallskip} \multicolumn{6}{c}{Fixed radius}\\ \multicolumn{6}{l}{Main sequence component}\\ $T_{\rm{eff}}$ (K) & 5931 & 5769 & 6095 & 5821 & 6035 \\ $\log{g}$ (dex) & 4.29 & 4.26 & 4.32 & 4.27 & 4.31 \\ E(B-V) (mag) & 0.001 & 0.000 & 0.012 & 0.000 & 0.008 \\ \\ \multicolumn{6}{l}{Subdwarf B component}\\ $T_{\rm{eff}}$ (K) & 33520 & 32400 & 34800 & 32764 & 34290 \\ $\log{g}$ (dex) & 5.81 & 5.77 & 5.85 & 5.79 & 5.83 \\ E(B-V) (mag) & 0.001 & 0.000 & 0.013 & 0.000 & 0.008 \\ \\ \multicolumn{6}{c}{Free radius}\\ \multicolumn{6}{l}{Main sequence component}\\ $T_{\rm{eff}}$ (K) & 5970 & 5730 & 6190 & 5840 & 6130 \\ $\log{g}$ (dex) & 4.38 & \multicolumn{2}{c}{/} & 3.55 & \,/ \\ E(B-V) (mag) & 0.001 & 0.000 & 0.100 & 0.000 & 0.090 \\ \\ \multicolumn{6}{l}{Subdwarf B component}\\ $T_{\rm{eff}}$ (K) & 33810 & 27500 & 45100 & 29000 & 40100 \\ $\log{g}$ (dex) & 5.90 & 4.70 & \,/ & 5.25 & \,/ \\ E(B-V) (mag) & 0.001 & 0.000 & 0.100 & 0.000 & 0.090 \\ \hline \end{tabular} \end{table} \begin{figure} \centering \includegraphics{f_models_sdb.eps} \caption{SED models calculated for the best fit parameters of PG\,1104+243 as given in Table\,\ref{tb-sedresults}, but with varying surface gravity of the sdB component. From bottom to top, the models have a $\log{g}$ of respectively 6.00, 5.85, 5.81, 5.77 and 5.00 dex (cgs). The crosses show the integrated model photometry. The photometric observations with their error bars are given in red.} \label{fig-models_sdb} \end{figure} \section{Gravitational Redshift}\label{s-gr} In an sdB + MS binary the difference in surface gravity between the two components is substantial. The surface gravity of a star gives rise to a frequency shift in the emitted radiation, which is known as the gravitational redshift. General relativity shows that the gravitational redshift as a function of the mass and surface gravity of the star is given by \citep{Einstein16}: \begin{equation} z_g = \frac{1}{c^2} \sqrt{G M g}. \label{e-gr} \end{equation} Where $z_g$ is the gravitational redshift, $c$ the speed of light, $G$ the gravitational constant, $M$ the mass, and $g$ the surface gravity. This $z_g$ will effectively change the apparent systemic velocity for the star. In a binary system the difference in surface gravity for both components will be visible as a difference in systemic velocity between both components. As the $z_g$ is proportional to the square root of the surface gravity, this effect is only substantial when there is a large difference in $\log{g}$ between both components, as is the case for compact subdwarfs and main-sequence stars. The measured difference in systemic velocity can be used to derive an estimate of the surface gravity of the sdB component. Using the mass ratio from the spectroscopic orbit, and the canonical mass of the sdB component, the mass of the MS component can be calculated. Combined with the surface gravity of the MS component from the SED fit, the $z_g$ of the MS component can be calculated. The $z_g$ of the sdB component can be obtained by combining the $z_g$ of the MS component with the measured difference in systemic velocity, and can then be converted to an estimated surface gravity of the sdB component. Using a canonical value of 0.47 $M_{\odot}$ for the sdB component, and the mass ratio derived in Sect.\,\ref{s-orbitalparamters}, we find a mass of 0.74 $\pm$ 0.07 $M_{\odot}$ for the MS component. The SED fit resulted in a $\log{g}_{\rm{MS}}$ of 4.29 $\pm$ 0.05 dex. Using equation \ref{e-gr}, the gravitational redshift of the MS component is $cz_g$ = 0.46 $\pm$ 0.05 km s$^{-1}$. The measured difference in systemic velocity is 1.97 $\pm$ 0.18 km s$^{-1}$, resulting in a total gravitational redshift of $cz_g$ = 2.43 $\pm$ 0.20 km s$^{-1}$ for the sdB component. This is equivalent to a surface gravity of $\log{g}_{\rm{sdB}}$ = 5.90 $\pm$ 0.08 dex, which is very close to the results of the SED fit. Other phenomena that can affect the presumed redshift of the MS component, as for example convectional blueshift (see, e.g., \citealt{Dravins82, Takeda12}), are not taken into account as the contribution of such effects is predicted to be around the one-sigma level. \section{Absolute dimensions}\label{s-absolutedimensions} Combining the results from the SED fit with the orbital parameters derived from the radial velocities, the absolute dimensions of PG\,1104+243 can be determined. With an assumed mass for the sdB component, the inclination of the system can be derived from the reduced mass determined in Sect. \ref{s-orbitalparamters}. Using this inclination, the semi-major axis of the system can be calculated. The surface gravity of the sdB component as determined by the SED fit and the gravitational redshift correspond within errors. For the final surface gravity, the average of both values is taken. The systemic velocity of PG\,1104+243 is determined by correcting the systemic velocity of the MS component with its gravitational redshift, resulting in $\gamma$ = $-$15.17 $\pm$ 0.07 km s$^{-1}$. The radius of both components is derived from their mass and surface gravity. The absolute dimensions of PG\,1104+243 are summarized in Table\,\ref{tb-absolutedim} For an sdB star of the canonical mass and a surface gravity of $\log{g}$ = 5.85 dex a temperature around 33000 K is expected, consistent with the result of the SED fit and the strength of the \ion{He}{i}/\ion{He}{ii} lines. This high surface gravity and temperature is consistent with the binary population models of \citet{Brown08} that find that low metalicity produces sdB stars located at the high $\log{g}$ end of the extreme horizontal branch. The canonical sdB mass assumption together with the derived mass ratio implies that the MS component must have a sub-solar metalicity. At solar metalicity it should have a radius of 0.8 $R_{\odot}$ and a temperature of $\sim$4800K, which is clearly ruled out by the SED fit. If we drop the canonical sdB mass assumption, and force the MS star to have normal metalicity, both the MS and sdB star must be more massive in order to be consistent with the temperature of the MS component and the spectroscopic mass ratio. From the He-MS models of \citep{Paczynski71}, one finds that a core-He burning model with M $>$ 0.6 $M_{\odot}$ would have Teff $>$ 40000K, which is ruled out by both SED fit and spectroscopy. Recent evolutionary models by \citet{Bertelli08} for main sequence stars with various metalicities, are perfectly consistent with the parameters from the SED fit for an 0.82 $M_{\odot}$ model, provided that the metalicity is assumed to be substantially sub-solar, at around $Z$ = 0.005. Such tracks are shown in Fig.\,\ref{fig-tracks}, where one can see that an 0.82 $M_{\odot}$ star has an expected temperature of around 5900 K, and a surface gravity of $\log{g}$ = 4.30 near the end of the main sequence, corresponding with an age of 15.0 $\pm$ 2.5 Gyr for the MS component. This is similar to the Hubble time and given all uncertainties in the models, indicates that the system is extremely old. The luminosity of both components can be calculated using $L = 4\pi \sigma R^2 T^4$, resulting in $L_{\rm{MS}}$ = 1.15 $\pm$ 0.13 $L_{\odot}$ and $L_{\rm{sdB}}$ = 22.5 $\pm$ 3.5 $L_{\odot}$. The apparent V magnitudes of both components are obtained directly from the SED fitting procedure and are given by: $V_{0,\rm{MS}}$ = 12.06 $\pm$ 0.05 mag and $V_{0,\rm{sdB}}$ = 12.03 $\pm$ 0.08 mag. The absolute magnitude can be obtained by integrating the best fit model SEDs over the Johnson V band, and scaling the resulting flux to a distance of 10 $pc$, resulting in $M_{V,\rm{MS}}$ = 4.64 $\pm$ 0.05 $mag$ and $M_{V,\rm{sdB}}$ = 4.61 $\pm$ 0.1 $mag$. The distance to the system can then be calculated from $\log{d} = (m_V − M_V + 5)/5$, which places the system at a distance of $d$ = 305 $\pm$ 10 pc. Obviously, the distance for both components comes out as exactly the same, as this is a fixed condition in the SED fitting process. The proper motion of PG\,1104+243 as measured by \citet{Hog00} is: \begin{equation} (\mu_{\alpha}, \mu_{\delta}) = (-63.4, -22.1) \pm (1.8, 1.9)\ \rm{mas\ yr^{-1}} \end{equation} Using the method of \citet{Johnson87}, these numbers together with the measured value of $\gamma$, can be used to compute the galactic space velocity vector \begin{equation} (U,V,W) = (-61.8, -55.3, -52.5) \pm (2.8,2.7,1.3)\ \rm{km\ s^{-1}} \end{equation} where U is defined as positive towards the galactic center. Using the values for the local standard of rest from \citet{Dehnen98}, we get \begin{equation} (U,V,W)_{\rm{LSR}} = (-51.8, -50.1, -45.3) \pm (2.8,2.8,1.4)\ \rm{km\ s^{-1}} \end{equation} Following the selection criteria of \citet{Reddy06}, PG\,1104+243 is bound to the galaxy, and belongs to the thick disk population. \begin{table} \centering \caption{Fundamental properties for both the main-sequence (MS) and subdwarf (sdB) component of PG\,1104+243.}\label{tb-absolutedim} \begin{tabular}{lrr} \hline\hline \noalign{\smallskip} \multicolumn{3}{l}{Systemic parameters} \\ $P$ (d) & \multicolumn{2}{c}{753 $\pm$ 3} \\ $T_0$ (HJD) & \multicolumn{2}{c}{2450386 $\pm$ 8} \\ $e$ & \multicolumn{2}{c}{$<$ 0.002} \\ $\gamma$ (km s$^{-1}$) & \multicolumn{2}{c}{$-$15.17 $\pm$ 0.07} \\ $q$ & \multicolumn{2}{c}{0.64 $\pm$ 0.02} \\ $a$ ($R_{\odot}$) & \multicolumn{2}{c}{322 $\pm$ 12} \\ $i$ $(^o)$ & \multicolumn{2}{c}{32 $\pm$ 3} \\ $E(B-V)$ & \multicolumn{2}{c}{0.001 $\pm$ 0.017} \\ $d$ (pc) & \multicolumn{2}{c}{305 $\pm$ 10} \\ \noalign{\smallskip} \multicolumn{3}{l}{Component parameters} \\ & \multicolumn{1}{c}{MS} & \multicolumn{1}{c}{sdB} \\ $K$ (km s$^{-1})$ & 4.42 $\pm$ 0.08 & 6.9 $\pm$ 0.3 \\ $M$ ($M_{\odot})$ & 0.74 $\pm$ 0.07 & \multicolumn{1}{c}{0.47 $\pm$ 0.05\tablefootmark{a}} \\ $\log{g}$ (cgs) & 4.29 $\pm$ 0.05 & 5.85 $\pm$ 0.08 \\ $R$ (R$_{\odot})$ & 1.02 $\pm$ 0.06 & 0.13 $\pm$ 0.02 \\ $T_{\textrm{eff}}$ (K) & 5930 $\pm$ 160 & 33500 $\pm$ 1200 \\ $L$ (L$_{\odot}$) & 1.15 $\pm$ 0.13 & 22.5 $\pm$ 3.5 \\ $V_0 $ (mag) & 12.06 $\pm$ 0.05 & 12.03 $\pm$ 0.08 \\ $M_V$ (mag) & 4.64 $\pm$ 0.05 & 4.61 $\pm$ 0.10 \\ \hline \end{tabular} \tablefoot{\tablefoottext{a}{Assumed value based on evolutionary scenario.}} \end{table} \begin{figure} \centering \includegraphics{f_tracks.eps} \caption{Evolutionary tracks calculated by \citet{Bertelli08} with a metalicity of $Z$ = 0.005, for three different masses: 0.60 $M_{\odot}$, 0.82 $M_{\odot}$ and 1.00 $M_{\odot}$. The results obtained for the cool companion of PG\,1104+243 are shown in red.} \label{fig-tracks} \end{figure} \section{Atmospheric parameters from \ion{Fe}{i-ii} lines}\label{s-ironlines} In this section the equivalent widths of iron lines are used to derive the atmospheric parameters and metalicity of the MS component. This method is independent of the sdB mass assumption used in the SED fit, and is used to confirm the effective temperature and surface gravity obtained with that method. Furthermore, the mass of the sdB component can be derived by using evolutionary tracks to fit the mass of the MS component, and the mass ratio from the radial velocity curves. Before the EWs can be measured, the continuum contribution of the sdB component must be subtracted from the HERMES spectra, which are then shifted to the zero velocity based on the MS radial velocities and summed. The LTE abundance calculation routine MOOG \citep{Sneden73} was used to determine the atmospheric parameters of the MS component using the EWs of 45 \ion{Fe}{i} and 10 \ion{Fe}{ii} lines in the wavelength range 4400--6850 \AA, excluding blended lines, and wavelength ranges contaminated by Balmer lines, \ion{He}{i} or \ion{He}{ii} lines from the sdB component. The atmospheric parameter determination also provides the metalicity of the MS component. The atomic data for the iron lines was taken from the VALD line lists \citep{Kupka99}, and the used oscillator strengths were laboratory values. The EWs are measured via integration and abundances are computed by an iterative process in MOOG in which for a given abundance, the theoretical EWs of single lines are computed and matched to the observed EWs. The effective temperature is derived by the assumption that the abundance of individual Fe I lines is independent of lower excitation potential, only Fe I lines are used because they cover a large range in lower excitation potential, resulting in $T_{\rm{eff}}$ = 6000 $\pm$ 250 K. A surface gravity of $\log{g}$ = 4.5 $\pm$ 0.5 dex is derived by assuming ionization balance between the iron abundance of individual \ion{Fe}{i} and \ion{Fe}{ii} lines. Assuming an independence between the iron abundance and the reduced EW, a microturbulence velocity of 2.0 $\pm$ 0.5 km/s can be derived. The final metalicity is based on the average abundance of all used lines, and result in a metalicity of [Fe/H] = -0.58 $\pm$ 0.11 dex, corresponding to $Z$ = 0.005 $\pm$ 0.001 using the conversion of \citet{Bertelli94}. The error on the metalicity is the line-to-line scatter for the Fe abundances for the derived atmospheric parameters. The used Fe I lines are fitted by creating synthetic spectra in MOOG using the derived atmospheric parameters, which confirm the obtained metalicity. A detailed description of the method used to determine the atmospheric parameters can be found in \citet[Sect. 3]{DeSmedt12}. These atmospheric parameters correspond very well with the effective temperature and surface gravity derived from the SED fitting process. The atmospheric parameters derived from the spectral analysis can be used to get a mass estimate for both the MS and sdB component independently from the SED fitting process, by using evolutionary models. Solar scaled evolutionary tracks of \citet{Bertelli08} with a metalicity of $Z$ = 0.005 are compared to the derived surface gravity ($\log{g}$ = 4.5 $\pm$ 0.5 dex) and effective temperature ($T_{\rm{eff}}$ = 6000 $\pm$ 250 K). The best fitting model has a mass of $M_{\mathrm{MS}}$ = 0.86 $\pm$ 0.15 $M_{\odot}$, with as corresponding age 12 $\pm$ 5 Gyr. Using the mass ratio from the radial velocity curves, a mass of $M_{\rm{sdB}}$ = 0.54 $\pm$ 0.10 $M_{\odot}$ is obtained for the sdB component. This corresponds within error to the canonical value of $M_{\mathrm{sdB}}$ = 0.47 $\pm$ 0.05 $M_{\odot}$. Using the newly obtained mass of the sdB component to derive its surface gravity from the gravitational redshift we find $\log{g}_{\rm{sdB}}$ = 5.84 $\pm$ 0.08 dex. Furthermore, the radius, luminosity and magnitudes in the V band can be derived using the methods described in Sect.\,\ref{s-absolutedimensions}, the results are shown in Table \ref{tb-absolutedim2}. Most of the derived parameters are compatible with those derived in Sect.\,\ref{s-absolutedimensions}. \begin{table} \centering \caption{Atmospheric parameters obtained based on the EW of iron lines, and derived properties. See Sect.\,\ref{s-ironlines} for further explanation.}\label{tb-absolutedim2} \begin{tabular}{lrr} \hline\hline \noalign{\smallskip} & \multicolumn{1}{c}{MS} & \multicolumn{1}{c}{sdB} \\\hline \noalign{\smallskip} $M$ (M$_{\odot})$ & 0.86 $\pm$ 0.15 & 0.54 $\pm$ 0.10 \\ $\log{g}$ (cgs) & 4.50 $\pm$ 0.50 & 5.84 $\pm$ 0.08 \\ $R$ (R$_{\odot})$ & 0.86 $\pm$ 0.07 & 0.15 $\pm$ 0.05 \\ $T_{\textrm{eff}}$ (K) & 6000 $\pm$ 250 & \multicolumn{1}{c}{/} \\ $L$ (L$_{\odot}$) & 0.87 $\pm$ 0.20 & \multicolumn{1}{c}{/} \\ $V_0 $ (mag) & 12.30 $\pm$ 0.15 & \multicolumn{1}{c}{/} \\ $M_V$ (mag) & 4.95 $\pm$ 0.15 & \multicolumn{1}{c}{/} \\ $d$ (pc) & 296 $\pm$ 12 & \multicolumn{1}{c}{/} \\ \hline \end{tabular} \end{table} \section{Discussion and Conclusions} From an analysis of literature photometry and observed spectra, detailed astrophysical parameters of PG\,1104+243 have been established. The surface gravity determined from the binary SED fitting method agrees very well with the surface gravity determined from the observed gravitational redshift. Furthermore, the long time-base observations made it possible to accurately establish the orbital period at 753 $\pm$ 3 d. The sdB component is consistent with a canonical post-core-helium-flash model with a mass of $\sim$0.47 $M_{\odot}$, formed through stable Roche lobe overflow. This is supported by the essentially circular orbit as stable mass transfer is the only known process that can circularize a long-period binary system. When comparing our results with the distribution predicted by BPS models, we find that the orbital period of PG\,1104+243 is higher than the most likely outcome at just over 100 days resulting from \citet[Fig. 21]{Han03}. However, the period is in agreement with the observed estimate of three to four years noted by \citet{Green01}, and with BPS models computed within the $\gamma$-formalism \citep{Nelemans10}. Our results therefore provide strong observational constraints for which BPS models are viable. \citet{Clausen12} published several limitations on the limiting mass ratio for dynamically stable mass transfer during the RLOF phase, the way that mass is lost to space and the transfer of orbital energy to the envelope during a common envelope phase, based on the observed period distribution of sdB + MS binaries. They state that if long-period (P$_{\rm{orb}}$ $>$ 75 d) sdB+MS binaries exist, mass transfer on the red giant branch (RGB) is stable, and if sdB + MS binaries with periods over 250 d exists, mass that leaves the system carries away a specific angular momentum proportional to that of the donor. In both cases the efficiency at which orbital energy is transferred to the envelope during the common envelope phase is below 75 \%. Furthermore, in several of their BPS models the resulting distributions for sdB + MS binaries would include periods up to 500 days \citep[Fig. 16]{Clausen12}. However, this formation channel cannot explain the low mass of the cool companion. In this formation channel, the cool companion would be expected to have a mass $\sim 1 M_{\odot}$. The low mass of the MS component (0.735 $\pm$ 0.07 $M_{\odot}$) indicates that during the stable Roche--Lobe overflow phase only a small amount of the mass was accreted, and the major part must have been lost to infinity. The high evolutionary age of the MS component ($\sim$13000 Myr) is further evidence of this. If the companion had accreted a significant fraction of the original envelope of the sdB component, its progenitor should have had a much lower mass as the current 0.735 $\pm$ 0.07 $M_{\odot}$. In this case, the MS component should be practically un-evolved, as e.g. the 0.6 $M_{\odot}$ track of Fig.\,\ref{fig-tracks} which hardly evolves at all in a Hubble time. As the time since the mass transfer ended cannot be longer than the core-helium burning lifetime of only ~100 Myr, no significant evolution of the MS component can have taken place since then. \citet{Deca12} published a study of the long-period sdB+K system PG\,1018--047, which shares many similar properties with PG\,1104+243. It has a period of 760 $\pm$ 6 d, a mass ratio of $M_{\rm{sdB}} / M_{\rm{K}}$ = 0.63 $\pm$ 0.11, and they find an effective temperature of 30500 $\pm$ 200 K and $\log{g}$ of 5.50 $\pm$ 0.02 dex for the sdB component. However, while the cool companion in PG\,1104+243 contributes $\sim$52\,\% of the light in the $V$-band, in PG\,1018--047 it only contributes $\sim$6\,\%, and is photometrically and spectroscopically consistent with a mid-K type star. \citet{Deca12} speculate that the orbit of PG\,1018-047 may be quite eccentric, and that it can have formed through the merger scenario of \citet{Clausen11}, meaning that the K-star was not involved in the evolution of the sdB star. The extremely low eccentricity of PG\,1104+243, makes such a scenario very unlikely, and we conclude that it is the first sdB+MS system that shows consistent evidence for being formed through post-Roche-lobe overflow. Thus, we have used high-resolution spectroscopy to solve the orbits of both components in an sdB+MS binary system, and have for the first time derived accurate and consistent physical parameters for both components. Furthermore, the accurate radial velocities allowed the first measurement of gravitational redshift in an sdB binary. PG\,1104+243 is part of an ongoing long-term observing program of sdB+MS binaries with HERMES at Mercator, and an analysis of the complete sample will make it possible to refine several essential parameters in the current formation channels. \begin{acknowledgements} We thank the referee for his useful suggestions. Many thanks to N. Cox, N. Gorlova and C. Waelkens for their help with obtaining spectra at the Mercator Telescope. Based on observations made with the Mercator Telescope, operated on the island of La Palma by the Flemish Community, at the Spanish Observatorio del Roque de los Muchachos of the Instituto de Astrofísica de Canarias. Based on observations obtained with the HERMES spectrograph, which is supported by the Fund for Scientific Research of Flanders (FWO), Belgium , the Research Council of K.U.Leuven, Belgium, the Fonds National Recherches Scientific (FNRS), Belgium, the Royal Observatory of Belgium, the Observatoire de Genève, Switzerland and the Thüringer Landessternwarte Tautenburg, Germany. Some of the observations reported in this paper were obtained at the MMT Observatory, a facility operated jointly by the University of Arizona and the Smithsonian Institution. The following Internet-based resources were used in research for this paper: the NASA Astrophysics Data System; the SIMBAD database and the VizieR service operated by CDS, Strasbourg, France; the ar$\chi$ive scientific paper preprint service operated by Cornell University. This publication makes use of data products from the Two Micron All Sky Survey, which is a joint project of the University of Massachusetts and the Infrared Processing and Analysis Center/California Institute of Technology, funded by the National Aeronautics and Space Administration and the National Science Foundation. The research leading to these results has received funding from the European Research Council under the European Community's Seventh Framework Programme (FP7/2007--2013)/ERC grant agreement N$^{\underline{\mathrm o}}$\,227224 ({\sc prosperity}), as well as from the Research Council of K.U.Leuven grant agreement GOA/2008/04, the German Aerospace Center (DLR) under grant agreement 05OR0806 and the Deutsche Forschungsgemeinschaft under grant agreement WE1312/41-1. P. Neyskens is Boursier FRIA, funded by the Fonds National Recherches Scientific. \end{acknowledgements} \bibliographystyle{aa}
\section{Introduction} For many years it has been believed that SU(N) lattice gauge theories exist in a single phase. Since confinement is a property of all compact lattice gauge theories at strong coupling, this assumption has the consequence that these theories also have a confining force (linear potential) in the weak coupling continuum limit. This fits rather nicely with phenomenological evidence that the strong interactions confine quarks. The numerical evidence for lack of a phase transition in SU(2) comes mostly from the apparently smooth behavior of the specific heat as a function of the coupling parameter, $\beta$. However, if the critical exponent $\nu$ is large enough to produce a large negative specific heat exponent $\alpha$ ($\alpha = 2-d\nu $), then the first infinite singularity may be in a high derivative, and not easily visible numerically. This is, for instance, often the case for 3-d spin glasses which are close to their lower critical dimension. The argument {\em in favor} of the existence a phase transition rests on the similarity of lattice gauge theories to ferromagnetic spin models, all of which in three or more dimensions, at least for short range interactions, have two phases, with the weak coupling phase exhibiting spontaneous symmetry breaking and long range order. Abelian and non-abelian spin models differ in detail, but {\em all} have phase transitions. A major difference between gauge models and spin models are that the former have local symmetries and the latter global. This difference can sometimes be erased through gauge fixing, however. For instance the 2-d lattice gauge theory is exactly equal to a set of non-interacting 1-d classical spin chains in the axial gauge, neglecting possible boundary effects. Setting all of the links in one direction equal to unity using the gauge freedom causes the 4-link plaquette action to collapse to a two-link dot product between the remaining gauge links which can be reinterpreted as spins.\cite{kogut} Another exact mapping is between the 3-d Ising gauge theory in axial gauge with the 3-d Ising model.\cite{wegner} For 4-d theories, the Coulomb gauge appears to be the most ``spin-like." Here the gauge freedom is used to maximize the traces of all of the links lying along three of the four lattice directions. In this gauge the fourth direction lying links become observables and act much like spins, in that they transform under a remnant 3-d {\em global} symmetry on each spacelike hyperlayer that remains after Coulomb gauge fixing. This symmetry can break spontaneously if the average of fourth-direction links over a hyperlayer acquires a magnetization. It has been shown that if this symmetry does break spontaneously, then the corresponding ferromagnetic phase is non-confining, so the average of these 4-th direction links serves as a local order parameter for confinement.\cite{zw,gs,mp} According to the standard hypothesis, the theory remains in the paramagnetic phase for all couplings in the non-abelian case, and undergoes a magnetic phase transition only in the abelian case, such as U(1). However, below it will be shown that this inconsistent with the known long-range order for classical 3-d Heisenberg models, to which the 4-d lattice gauge theories are closely related and actually connected in a larger coupling space. Numerical evidence for a ferromagnetic phase transition based on the behavior of the Coulomb gauge spin-like order parameter described above has been given for SU(2) with the Wilson action. Finite size scaling shows an infinite-lattice transition around $\beta = 3.2$ with a correlation length critical exponent $\nu = 1.7\pm 0.2$.\cite{cgm} Other simulations\cite{cgm} which supplement the Wilson action with an infinite chemical potential for gauge-invariant SO(3)-Z2 monopoles\cite{so3-z2}, together with a positive plaquette constraint, appear to remain in the ferromagnetic phase for all couplings. This would seem to indicate that the normal confinement seen with the Wilson action is due to strong-coupling lattice artifacts, not unlike the U(1) case. In the following, the analytic case for the existence of this phase transition is explored. An analytic proof for the opposite hypothesis, namely the lack of a phase transition with confinement persisting to the continuum limit, was presented by Tomboulis some time ago\cite{tomboulis} and updated more recently.\cite{tomboulisnew} This proof is fairly complex and some possible flaws have been noted.\cite{seiler} When an hypothesis is difficult to prove, it is sometimes worthwhile to attempt to prove the opposite, which is the approach taken below. \section{Extended Coupling Plane} The 4-d SU(2) theory can be extended into a larger coupling space by allowing the coupling for purely spacelike (horizontal) plaquettes to differ from that of plaquettes which include the Euclidean timelike direction (vertical). The action is \begin{eqnarray} S & = & \sum _{\vec{n}} \left(\beta _H \sum _{i=1}^{2} \sum _{j=i+1}^{3} (1-\frac{1}{2} \mathrm{tr}( U_{\vec{n},i} U_{\vec{n}+\hat{\imath},j} U_{\vec{n}+\hat{\jmath},i}^{\dagger}U_{\vec{n},j}^{\dagger}))\right. \\ & & +\mbox{}\beta _V \left. \sum _{i=1}^{3} (1-\frac{1}{2} \mathrm{tr}( U_{\vec{n},i}U_{\vec{n}+\hat{\imath},4} U_{\vec{n}+\hat{4},i}^{\dagger}U_{\vec{n},4}^{\dagger}))\right) \nonumber \end{eqnarray} Where the $U_{\vec{n},j}$ are SU(2) valued gauge links based at lattice site $\vec{n}$ in direction $j$. The normal Wilson action is simply the $\beta _V = \beta _H$ case, however because the renormalization group gives a relation between $\beta$ and the physical lattice spacing through the running coupling, the $\beta _V \ne \beta _H$ theory can also be considered an SU(2) LGT with unequal lattice spacings in the spacelike and Euclidean timelike directions. So at least in the phase connected to the continuum limit (neighborhood of ($\beta _ H$, $\beta _ V$) $\rightarrow$ ($\infty$,$\infty$)) the entire phase plane excluding the boundaries can be considered an SU(2) LGT. Consider now the $\beta _H = \infty$ theory. This can be seen to be equivalent to a set of non-interacting 3-d O(4) classical Heisenberg models as follows. At $\beta _H = \infty$ all of the horizontal plaquettes will be forced to their largest possible value of unity. One can then find a gauge in which all horizontal links are unity also, as follows. Set a maximal tree of links to unity on each spacelike hypersurface. This is a partial axial gauge in which the final trunk of the tree along the Euclidean timelike direction is not completed. For instance for an $L^4$ lattice, on a given spacelike hyperlayer all 3-direction links can be set to unity except when $z=0$. On the $z=0$ plane all 2-direction links are set to unity except for when $y=0$ and along the line ($y=0$, $z=0$) all 1-direction links except for when $x=0$ can be set to unity by gauge transformations. Looking at the $z=0$ plane, there is a set of $x$-$y$ plaquettes extending backward from $x=0$ (through the periodic boundary condition) which have three links set to unity. The plaquette being unity due to $\beta _H = \infty$ ensures that the fourth link in the plaquette is also unity. Now the same is true for the next row of plaquettes etc., forcing all links to unity except for the last row pointing in the positive direction from x=0. These are now equal to the gauge invariant Polyakov loop for that direction, and all of these links are equal. This same procedure can be extended from the ($x=0$, $y=0$) plane along the $z$ direction to show that the links out of the plane are also unity, except for one at the edge along each lattice direction which are equal to their neighboring links pointing in that direction. Finally, a Polyakov loop symmetry transformation (also a symmetry of the action) can be employed to bring these final set of links to unity. In this restricted sector, the Polyakov loop symmetry is the whole SU(2) group rather than just the center. One could also formulate the theory with open boundary conditions, in which case this final step would be unnecessary. Although the language of axial gauge was used above, one can see that actually the condition for Coulomb gauge has also been met, that the sum of traces of all horizontal links is maximized. Coulomb gauge leaves a layered remnant symmetry, one global SU(2) per hyperlayer, unfixed. Because Coulomb gauge was being sought is why the axial tree above was left uncompleted. Since all of the horizontal plaquettes are unity, the vertical plaquettes simplify to \begin{equation} \frac{1}{2}\mathrm{tr}(U_{\vec{n},4}^{\dagger}U_{\vec{n}+\hat{\jmath},4}) \end{equation} where $j$, of course, can be 1, 2 or 3. Writing each $U$ as \begin{equation} U=s_0 + i\sum _{k=1}^{3} s_k \tau _k \end{equation} where the $\tau _k$ are the Pauli matrices, one can associate an O(4) unit vector $\vec{s}=(s_0$, $s_1$, $s_2$, $s_3)$ with the SU(2) valued link. In terms of the $\vec{s}$ 's it is easily verified that the vertical plaquette above is simply the nearest neighbor O(4) dot product, \begin{equation} \frac{1}{2}\mathrm{tr}(U_{\vec{n},4}^{\dagger}U_{\vec{n}+\hat{\jmath},4})=\vec{s}_{\vec{n}}\cdot \vec{s}_{\vec{n}+\hat{\jmath}} \end{equation} The action on each hyperlayer is that of the 3-d O(4) Heisenberg model with coupling parameter (inverse temperature) $\beta _V$. In a gauge theory there are never any direct interactions between links longitudinally, and the freezing of horizontal links prevents any indirect interactions among hyperlayers, so these all become independent Heisenberg models. Each of these has its own SU(2) global symmetry, the remnant symmetry from the Coulomb gauge. The reason behind this is that an SU(2) gauge transformation which is {\em global} on the hyperlayer transforms horizontal links in such a way as to leave the trace of each link unchanged. Thus the gauge condition, which is to maximize these traces, is not disturbed. Therefore, at the level of the partition function, the $\beta _H = \infty$ SU(2) lattice gauge theory becomes a set of non-interacting 3-d O(4) classical Heisenberg Models. Durhus and Fr\"{o}hlich\cite{df} earlier pointed out a connection between SU(2) lattice gauge theory and the 3-d O(4) model, but did not consider the case of split vertical and horizontal couplings which allows for an exact mapping. Long range order (LRO) in the classical Heisenberg model has been rigorously proven.\cite{lro} This means that a ferromagnetic phase must exist at a finite weak coupling (large $\beta _V$). Because this is a symmetry broken phase, it must be separated from the strong-coupling paramagnetic phase by a phase transition, which has been convincingly found by Monte Carlo simulation at a coupling $\beta = 0.9360(1)$.\cite{o4} The LRO proof does not depend on the specific symmetry group. Indeed, the existence of phase transitions in all ferromagnetic spin models is well-established and non-controversial from both analytic and numerical perspectives. The ferromagnetic order for the Heisenberg model is also quite robust to the addition other interactions. For instance it is still preserved at zero temperature even if up to 20\% of the interactions are switched to anti-ferromagnetic (beyond this level it enters a spin glass phase)\cite{spinglass}, and at higher temperatures with smaller but still substantial contaminations. Below, it will be argued that the ferromagnetic phase, and therefore also the phase transition, continues to exist for non-infinite $\beta _H$ as well. In Fig.~1 the $\beta _H$ - $\beta _V$ phase plane is shown, with the 3-d O(4) model existing on the top edge ($\beta _H = \infty$). The normal 4-d SU(2) model exists on the $\beta _V =\beta _H$ line. However, as stated before, the $\beta _V \ne \beta _H$ cases can also be considered to be SU(2) lattice gauge theories, with a different lattice spacing in the fourth direction than the other three. Thus the entire interior, at least in the vicinity of the continuum limit in the upper right corner, is an SU(2) lattice gauge theory. \begin{figure}[ht] \includegraphics[width=3.5in]{lrofig1.eps} \caption{Possible phase diagram for SU(2) lattice gauge theory with different horizontal and vertical couplings. Top horizontal axis is 3-d O(4) Heisenberg model and the star shows its known ferromagnetic phase transition.\protect\cite{o4} Square and triangle show locations of phase transitions at $\beta _H = 20$ and $\beta _ H = \beta _V$ found from Monte Carlo simulations using the Coulomb gauge magnetization.\protect\cite{cgm} Diamond is possible spin-glass to paramagnetic transition found previously using the real replica method.\protect\cite{su2sg} Bold regions of upper and right edges show portions of the border known to be ferromagnetic. Lines drawn are hypothetical phase boundaries guided by the Monte Carlo results.\protect\label{fig1}} \end{figure} The bottom edge, with $\beta _H =0$, has only the vertical plaquettes left. In an axial gauge with the fourth direction links set to unity, the system becomes a set of 1-d spin chains, which are paramagnetic except at zero temperature ($\beta _V = \infty$) where there is a phase transition to the ferromagnetic ground state. The right edge of the phase diagram, $\beta _V = \infty$, is an odd phase where all of the vertical plaquettes are forced to unity. Staying with the axial gauge, one can see that the 3-d SU(2) theories on spacelike hypersurfaces but different Euclidean times are {\em locked} to each other, in other words they can only fluctuate in lockstep. This makes for a theory with a 4-d energy, but non-extensive 3-d entropy, so it is basically stuck in the classical ground state. Both top and bottom ends of this line are ferromagnetic; there is no reason for the whole line not to be also. Thus ferromagnetism exists on the upper border for $\beta _V > 0.936$ and along the entire right-side border. In the following, two arguments are given which show that the ferromagnetic phase persists as one enters the interior of the phase diagram. An important point is that the global SU(2) symmetries of the independent Heisenberg models persist for the $\beta _H < \infty$ case in the form of the Coulomb Gauge remnant symmetries which exist independently on each hyperlayer. For a symmetry breaking phase transition, once the phase-transition line has entered the phase plane it must continue to another edge. This is because the order parameter, due to the realized symmetry, is exactly zero in the paramagnetic region on the infinite lattice, and of course is nonzero in the ferromagnetic phase. An analytic function zero in a finite region is zero everywhere, so a line of non-analyticity must separate the two phases completely. The only place where the phase transition line could terminate is at the only other phase transition on the border, namely the lower right corner as shown. Because the remnant symmetry exists for both the $\beta _H = \infty$ and the $\beta _H$ finite cases, this situation is not analogous to the Ising model in an external field, where the field, no matter how small, removes the transition due to explicit symmetry breaking. Fig.~1 also shows the ferromagnetic phase transitions found with Monte Carlo simulations using the Coulomb Gauge magnetization as the order parameter. One of these was for $\beta _H = 20$ where the phase transition on the infinite lattice was determined to be $\beta _{Vc} = 1.01(2)$. This transition closely resembles that of the Heisenberg model in that it has similar critical exponents. Finite size scaling fits are good, which indicates lattices are large enough to suppress non-leading effects. Observation of such a transition in Monte Carlo simulations is a strong indication that the phase transition does enter the interior of the phase diagram. Below, it will be argued that that must be the case from an analytic perspective as well. \section{Persistence of Phase Transition for $\beta _H < \infty$} In order for the SU(2) lattice gauge theory in the interior of the phase diagram to avoid the phase transition as in the conventional hypothesis, the transition would need to be somehow prevented from entering the interior of the phase diagram at all. This means that the small terms that arise in the Hamiltonian when $\beta _H$ is backed off from $\infty$ would have to destroy the ferromagnetic order, no matter how small the coupling $g_H$ ($\beta _H = 4/g_{H}^{2}$). This seems odd considering the continued existence of the symmetry and the known robustness of ferromagnetism in the Heisenberg model. Say we start in the deep-ferromagnetic region of the Heisenberg model. For the order parameter to jump from a finite value at $g_H =0$ to a value of zero for any $g_H >0$ there would have to be a first-order phase transition at the edge of the phase diagram. However, this edge is already well characterized, since it is the ferromagnetic phase of the Heisenberg model itself. Known properties of this phase are inconsistent with it being the location of a first-order phase transition in the {\em same order parameter} due to the following. A symmetry-breaking first-order phase transition is of the type associated with a tricritical point, which is controlled by a sixth-order Landau effective potential as shown in Fig.~2 (solid line) at the point of phase transition.\cite{cl} For $g_H >0$ the two side-minima would hypothetically lift up, leaving only the minimum at zero order parameter. On the phase transition point, both the phase with zero order parameter {\em and} the two instances of the broken-symmetry phase exist in equilibrium. Such phase mixing would be easily observable through an un-sharp order parameter and a latent heat (range of internal energies). However, such a state simply does not exist in the Heisenberg model. In the deep ferromagnetic region its Landau effective potential is widely believed to look like the dashed line in Fig.~2 with only two minima which is incompatible with the tricritical behavior. \begin{figure}[ht] \includegraphics[width=3.5in]{lrofig2.eps} \caption{Landau effective potentials at a symmetry breaking first-order transition (solid line) and in the deep-ferromagnetic region of a system with a higher-order transition (dotted line). Scales are arbitrary.} \label{fig2} \end{figure} Therefore, the hypothesis that the ferromagnetic phase does not enter the interior of the phase diagram seems to require behavior of the Heisenberg model inconsistent with known behavior. If a first-order phase transition is present it cannot be a conventional one. However, the argument above is somewhat heuristic and non-rigorous in that it involves Landau effective potentials. A more rigorous proof of the impossibility of a first-order phase transition as one enters the phase plane from the deep-ferromagnetic region of the Heisenberg model can be constructed as follows. Consider the correlation function $C(\vec{q},\vec{r})=<\frac{1}{2}\mathrm{tr}(U_{\vec{q},4}U_{\vec{r},4}^{\dagger})>$. In the ferromagnetic phase this shows spontaneous symmetry breaking through \begin{equation} \lim _{|\vec{q}-\vec{r}|\rightarrow \infty} C(\vec{q},\vec{r})=<|\vec{m}|>^2 \end{equation} where \begin{equation} \vec{m}=\frac{1}{V}\sum _{\vec{n}}\vec{s}_{\vec{n}} \end{equation} where the sum is over individual 3-d hypersurfaces and $V$ is the 3-volume. The expectation value above includes an average over different Euclidean times as well as gauge configurations. Although a finite lattice is used for defining quantities, the infinite lattice case is implicitly being considered here. Now let us calculate the derivative of this correlation function with respect to $g _H$ at $g _H = 0$. Note that the calculation is performed in the Heisenberg model itself (earlier shown to be the $g _H = 0$ limit), but in order to determine what function to calculate the expectation value of, we must consider the $g _H \ne 0$ theory. To do this, rewrite the horizontal $U$ links in terms of gauge fields $A$, as in the usual calculation of the continuum limit: \begin{eqnarray} S & = & \sum _{\vec{n}}\left( \frac{4}{g_{H}^{2}}\sum _{i,j>i}(1-\frac{1}{2}\mathrm{tr}( \exp (ig_H \vec{A}_{\vec{n},i}\cdot \vec{\tau}) \exp (ig_H \vec{A}_{\vec{n}+\hat{\imath},j}\cdot \vec{\tau})\right.\nonumber\\ & & \;\;\;\;\;\;\;\;\;\;\;\; \exp (-ig_H \vec{A}_{\vec{n}+\hat{\jmath},i}\cdot \vec{\tau}) \exp (-ig_H \vec{A}_{\vec{n},j}\cdot \vec{\tau}))) \nonumber\\ & & \!\!\!\! +\mbox{}\beta _V \! \! \left. \sum _{i<4} (\! 1\! -\! \frac{1}{2}\mathrm{tr}( \exp (ig_H \vec{A}_{\vec{n},i}\cdot \! \vec{\tau}) U_{\vec{n}+\hat{\imath},4} \exp (-ig_H \vec{A}_{\vec{n}+\hat{4},i}\cdot \! \vec{\tau}) U_{\vec{n},4}^{\dagger}))\! \! \right) \end{eqnarray} where $i$,$j$ run from 1 to 3. No approximation has been made. For finite $g _H$ this is not a practical decomposition, since the A integrations still need to obey the Harr measure for the $U$'s, however it is useful in the limit $g _H \rightarrow 0$. Taking the limit $g _H \rightarrow 0$ gives \begin{equation} \label{action0} S=\sum _{\vec{n}}\left( \sum _{i,j>i}F_{ij}^2+\beta _V \sum _{i} (1-\frac{1}{2}\mathrm{tr}(U_{\vec{n}+\hat{\imath},4} U_{\vec{n},4}^{\dagger}))\right) \end{equation} where $F_{ij}$ is the abelian field strength tensor (the non-abelian part having a factor of $g _H $). This gives the multiple 3-d Heisenberg models as expected, but also a disconnected set of 3-d free field theories. The connected part of \begin{equation} \left. \frac{\partial C(\vec{q},\vec{r})}{\partial g_ H }\right| _{g _H =0} \end{equation} is given by \begin{equation} \frac{1}{4}\beta _V \sum _{\vec{n},i<4}<\mathrm{tr}(U_{\vec{q},4}U_{\vec{r},4}^{\dagger})\mathrm{tr}((i\vec{A}_{\vec{n},i}\cdot \vec{\tau}) U_{\vec{n}+\hat{\imath},4}U_{\vec{n},4}^{\dagger})>. \end{equation} and a similar term with the $\vec{A}$ in the other position. This is to be computed in the $g_H = 0$ action of Eq. \ref{action0} which is even in the transformation $\vec{A}\rightarrow -\vec{A}$. As a consequence, any expectation value containing a single factor of $\vec{A}$ is zero. Disconnected parts similarly vanish or cancel. This almost trivial observation means that the first derivative of the order parameter with respect to $g_V$ at $g_V=0$ vanishes, and therefore exists. If the first derivative exists at a point, it follows from an elementary theorem of analysis that the function itself is continuous at that point.\cite{rudin} However, if the order parameter is a continuous function of $g_V$ at $g_V =0$ then it cannot drop suddenly to zero here in a first-order phase transition, which would require a discontinuity. Therefore, the ferromagnetic phase must enter the phase diagram. In fact, since $g _H$ does not affect the phase transition at lowest order, the line would be expected to enter at a $90 ^{\circ}$ angle. Since, as argued before, the entire interior of the phase diagram is an SU(2) LGT if unequal lattice spacings are considered, this means that there must be a ferromagnetic phase in SU(2) as well. The SU(2) LGT has long-range order. The order parameter is the Coulomb-gauge magnetization. As argued before, due to the symmetry breaking nature of the phase transition, the phase transition line must continue to another edge, the only possibility being the lower right corner of Fig.~1. In doing so it clearly must cross the $\beta _V = \beta _H$ line, which is the ordinary SU(2) LGT with equal lattice spacings. A phase transition in which the Coulomb Gauge remnant symmetry breaks spontaneously is known to be deconfining.\cite{zw,gs,mp} Therefore the zero physical temperature (infinite lattice) 4-d SU(2) lattice gauge theory must have a deconfining phase transition, contrary to the usual assumption. To prove for SU(3) or SU(N) one only has to replace the O(4) Heisenberg model with the SU(N)$\times$SU(N) spin model. Since these all have ferromagnetic phases, the argument goes through in the same way for them. This argument has similarities to another approach which had the same conclusion.\cite{fa} In the fundamental-adjoint plane of SU(2) lattice gauge theory with couplings $\beta _F$ and $\beta _A$, there is a well-known first-order phase transition which starts at the 4-d Z2 LGT transition (at $\beta _A = \infty$) and ends in the middle of the diagram at ($\beta _F$, $\beta _A$) = (1.48, 0.90).\cite{bc} This has been seen as a critical point, below which an analytic path exists between the strong-coupling confining region with the weak-coupling continuum limit. However, that would require the first-order phase transition to be a non-symmetry-breaking transition. If it were symmetry-breaking, on the other hand, then the end of the first-order line would be a tricritical point, and the transition would have to continue as a higher-order one, bisecting the entire coupling plane into symmetry-broken and unbroken sectors, as above. Interestingly, a critical point and a tricritical point are rather easily distinguished through the scaling behavior of the attached first-order transition. In particular, the scaling relationship between the latent heat and the size of the metastability region is linear in the critical case and quadratic for the tricritical case.\cite{fa} One simply monitors the shape of the growing hysteresis rectangle in the energy-coupling plane while moving up the first-order line. The numerical evidence points convincingly to the tricritical case.\cite{fa} This again implies a symmetry-breaking phase transition separates the strong and weak-coupling regions. One of the strengths of this energy-scaling argument is that it does not require identification of the symmetry or order parameter involved, and no gauge-fixing is required. The tricritical case simply requires {\em some} symmetry to break. However, it seems likely it is the same symmetry breaking as studied above, something relatively easy to check. \section{Artifact Driven Transition} It is interesting to consider possible mechanisms which could drive the SU(2) transition. In the U(1) case, confinement arises from abelian monopoles, which are strong-coupling lattice artifacts.\cite{dgts} The confined phase occurs when the monopoles form percolating loops. Confinement can be prevented by suppressing monopoles with a chemical potential.\cite{mitrj} Some time ago a gauge invariant monopole was introduced which could play the same role for the SU(2) theory.\cite{so3-z2} It carries SO(3) and Z2 degrees of freedom which in some sense cancel each other, so was named an SO(3)-Z2 monopole. Suppressing these monopoles, together with a positive plaquette restriction, appears to prevent the transition to the confining phase.\cite{cgm} The system stays in the spontaneously magnetized phase all the way to zero coupling when this constraint is imposed.\cite{toap} Lattices to $60^4$ have been measured and the lattice spacing (determined from the running coupling) is such that these should definitely be in the confining region if universality applies. Potentials, measured to 1.5 fm, show no evidence of a linear term. \begin{figure}[ht] \includegraphics[width=3.5in]{lrofig3.eps} \caption{Percolation probabilities for SO(3)-Z2 monopoles as a function of $\beta$ for various lattice sizes (diamond $16^4$, * $18^4$, $\times$ $20^4$, box $24^4$, triangle $30^4$, + $40^4$). Also shown is extrapolation of 50\% point to the infinite lattice (uses right y-axis).} \label{fig3} \end{figure} It is interesting to monitor these monopoles in the standard Wilson theory on periodic lattices. They form percolating clusters in the crossover region where confinement occurs and de-percolate at weaker couplings. Fig.~3 shows the percentages of percolating lattices as a function of $\beta$ for various lattice sizes. The finite-lattice percolation transition can be taken to occur at the 50\% level. Also shown is the extrapolation of the percolation transition to the infinite lattice. The extrapolation shown uses the finite-lattice shift equation for a higher-order phase transition\cite{barber} \begin{equation} 1/\beta _{cL}= 1/\beta _{c}+cL^{-1/\nu} \end{equation} using the expected value $\nu = 1.7$ from Ref.~6. Here $L$ is the linear lattice size and $\beta _{cL}$ is the apparent critical point on the finite lattice. A fit to the data for $c$ and $\beta _c$ gives $\beta _c = 3.161(2)$. If instead, $\nu$ is assumed to be unknown and determined from the fit then the best fit is obtained for $\nu = 1.1$ yielding $\beta _c = 3.193(1)$. Setting $\nu = 0.5$ gives $\beta _c = 3.226(2)$. The uncertainties in $\beta_c$ from the fixed-$\nu$ fits are much smaller than the differences in fits with different $\nu$'s, so the majority of uncertainty in $\beta _c$ is from the extrapolation. Allowing a broad range for this exponent from 0.5 to 1.7 gives for the infinite lattice $\beta _c = 3.19(3)$. It is very interesting that this is consistent with the position of the deconfining Coulomb-gauge magnetization transition determined from data-collapse fits to scaling behavior, which yielded an infinite lattice critical point of $3.18(8)$.\cite{cgm} It seems unlikely that these agree by mere coincidence. The percolation study was done after the other study was completed and released. Being that the percolation study was performed on the standard Wilson theory using conventional heat-bath and over-relaxation updates and with periodic boundary conditions, the coincidence of these results for $\beta _c$ gives additional confidence that the open boundary-condition Coulomb-gauge methods used in Ref.~6 are reliable. The fact that the percolation transition moves to {\em smaller} $\beta$ as the lattice size is increased means it almost certainly exists on the infinite lattice. The connection between SO(3)-Z2 monopole percolation and confinement will be further explored in a forthcoming publication\cite{toap}. \section {Conclusion} It has long been known that lattice gauge theories have a lot in common with spin models. This becomes especially apparent in certain fixed gauges for which the remnant symmetry can be matched onto a spin model. The result presented above implies spin and gauge theories have even more in common, namely long range ferromagnetic order. This has important consequences for the non-abelian lattice gauge theories SU(2) and SU(3) which were previously assumed to be exceptions to this behavior. The relatively large critical exponent, $\nu \sim 1.7$ for SU(2), explains how such a transition could have been missed, since the corresponding specific heat singularity is very soft. The ordered continuum limit means that the quenched lattice gauge theory without fermions does not confine in the continuum limit, an idea which has been suggested previously.\cite{previous,ps} Because good phenomenological evidence exists for quark confinement, however, some source of confinement must be sought, which need not necessarily be a linear potential. If light quarks are added to the theory, chiral symmetry is still expected to break spontaneously through the formation of a quark condensate. This only requires a sufficiently strong force, not necessarily a confining one. If this collective state polarizes in such a way as to repel strong color fields, the chiral condensate could form a kind of bag surrounding mesons and baryons, contributing a confining-like term to the force over a limited range, though confinement would not be absolute.\cite{csbqc,ch,gribov} Also, if a quark were to find itself a long distance from its partner antiquark, a quark/antiquark current in the chiral condensate could quickly generate local partners. This is somewhat different from generating quark anti-quark pairs from gluon ``sparking" which originate at the same location, but that too can prevent isolated quarks from existing provided there is sufficient energy in the bond. Thermodynamics of the quark-gluon plasma is also modified in this picture, since confinement is no longer in the gluon sector. A phase transition or at least a rapid crossover to a chiral-symmetric phase at high physical temperatures will undoubtedly still exist, which, if chiral symmetry breaking is related to confinement as above, will also be deconfining.
\section{Introduction} Resistive Plate Counters\cite{Santonico:1981sc} (RPC) detectors are installed at both the ATLAS (A Toroidal LHC Apparatus)\cite{atlas} and CMS (Compact Muon Solenoid)\cite{Chatrchyan:2008aa} experiments at the LHC (Large Hadron Collider) of CERN, Geneva (Switzerland) to provide triggering and synchronization in both barrel and endcap regions as part of the muon system. RPCs use a freon-based gas mixture (typically 95.2\% C$_2$H$_2$F$_4$ - 4.5\% Iso-C$_4$H$_{10}$ - 0.3\% SF$_6$) in a recirculation system call Closed Loop (CL)\cite{gassystem}. Gas mixture is humidified at the 40\% RH level to balance the ambient humidity that affects the resistivity of the highly hygroscopic bakelite. The CL was designed to cope with large gas mixture volumes and costs. In the closed loop system industrial filters commercially available are in operation to purify the mixture and to prevent contamination collection that affects the RPC performances. \par A systematic study of CL gas purifiers has been carried out from 2008 to 2011 at CERN using RPC chambers exposed to cosmic rays and a scaled-down closed loop gas system equipped with several gas analysis sampling points. Goals of the study\cite{Abbrescia:2006LNF} were to observe the release of contaminants in correlation with the dark current increase in RPC detectors, to measure the purifier lifetime\cite{Benussi:2010yx} with unused material, to observe the presence of pollutants. In this paper, new preliminary results from the 2011 run are shown which characterize the behavior of used purifiers and study the pattern of dark currents increase in the upstream versus the downstream gaps. \section{Experimental setup} The experimental setup\cite{Bianco:2009CMSNOTE}\cite{colafranceschi_thesis} is composed of a closed and an open loop gas systems (Fig.~\ref{FIG:isr_setup}). To purify the closed loop gas mixture (that continuously receives 10\% of fresh mixture), commercial filters are used as shown in Fig.~\ref{FIG:chem_setup}. \par The first purifier consists of a 5\AA~~ (10\%) and 3\AA~~ (90\%) type~ zeolite molecular sieve (ZEOCHEM\cite{ZEOCHEM}). The second purifier cartridge is filled with 50\% Cu-Zn filter type R12 (BASF\cite{BASF}) and 50\% Cu filter type R3-11G (BASF) while the third purifier consists of NiAlO$_3$ filter type 6525 (LEUNA\cite{LEUNA}). \begin{figure}[h] \begin{center} \resizebox{11.0cm}{!}{\includegraphics{setup_bn.eps}} \caption{Schema of the CL setup.} \label{FIG:isr_setup} \end{center} \end{figure} \begin{figure}[h] \begin{center} \resizebox{9.7cm}{!}{\includegraphics{sampling_point.eps}} \caption{Purifiers equipped with several sampling points (HV61, 62, 64) before and after each stage.} \label{FIG:chem_setup} \end{center} \end{figure} Eleven double-gap RPC detectors are installed in a temperature (Fig.~\ref{FIG:isrtemp}) and humidity (Fig.~\ref{FIG:isrrh}) controlled hut, with online monitoring of environmental parameters. \begin{figure}[h] \begin{center} \resizebox{11.7cm}{!}{\includegraphics{ISRtemp.eps}} \caption{Temperature trend inside and outside the experimental hut.} \label{FIG:isrtemp} \end{center} \end{figure} \begin{figure}[h] \begin{center} \resizebox{11.7cm}{!}{\includegraphics{ISRrh.eps}} \caption{Relative humidity trend inside and outside the experimental hut.} \label{FIG:isrrh} \end{center} \end{figure} Nine detectors out of eleven are operated in CL mode while two are operated in open loop (OL) mode. Each RPC detector has two gaps (upstream and downstream) whose gas lines are serially connected. The detectors are operated at a 9.2~kV voltage supply. At the working point selected, the anode dark current drawn (due to the high bakelite resistivity) is approximately 1-2~$\mu$A. Gas sampling points, before and after each filter of the closed loop, allow chemical and gaschromatograph (GC) analysis\cite{Bianco:2009CMSNOTE}. Gas mixture composition is monitored twice a day by GC, which also provides the amount of air contamination, stable over the entire data-taking run and below 300 (100) ppm in closed (open) loop as shown in Fig.~\ref{air_cont}. \begin{figure}[h] \begin{center} \resizebox{11cm}{!}{\includegraphics{air_contamination.eps}} \caption{Air contamination measured in the open and closed loop recirculation system.} \label{air_cont} \end{center} \end{figure} \section{Chemical analysis setup} Chemical analyses have been performed in order to correlate the increase of dark currents with the release of gas contaminants. To identify the contaminants nature, the gas is sampled before and after each purifier, and bubbled into a set of PVC (Polyvinyl chloride) flasks (Fig.~\ref{FIG:sampling_point}). \begin{figure}[h] \begin{center} \resizebox{9.0cm}{!}{\includegraphics{chem_setup.eps}} \caption{Chemical analyses are performed using LiOH flasks in which gas is bubbled and contaminants collected.} \label{FIG:sampling_point} \end{center} \end{figure} Flask 1 acts as a buffer to avoid return of LiOH into the CL. Flasks 2, 3 and 4 contained 250~ml solution of LiOH (0.001~mol/l corresponding to 0.024~g/l, optimized to keep the pH of the solution at 11). The bubbling of gas mixture into the three flasks allows one to capture a wide range of elements that are likely to be released by the system, such as Ca, Na, K, Cu, Zn, Ni, F. At the end of each sampling line the flow is measured to estimate the total amount of gas for the whole period of data-taking. Sampling points HV61 and HV64 (Fig.~\ref{FIG:sampling_point}) are located before filters (HV61), after zeolite filter (HV62), after Cu/Zn filter (HV64), after Ni filter (HV66). \par The fluorine production of RPCs in CL was measured previously in high-radiation condition \cite{Aielli05},\cite{Abbrescia:2006hk}. To measure the fluorine production, sampling point HV61 and HV62 are equipped with two additional flaks and fluoride selective electrodes. The $\rm{F^-}$ selective electrode adopted\cite{Hanna} is a solid state half-cell sensor that requires, as a separate reference, a silver-silver chloride double junction half-cell reference electrode. Selective electrodes are installed to measure the ionic potential, which is directly connected to the ionic concentration. The selective electrodes monitor the collected amount of ions integrated over time by means of a custom software which logs the electrochemical potential every 10 minutes in order to estimate the $\rm{F^-}$ production rate and concentration in the system. Both electrodes (reference and sensor) are immersed in a diluted TISAB II solution. This increases and stabilizes the ionic strength of the solution making a linear correlation between the logarithm of the concentration of analyte and the measured potential. The selective electrodes were calibrated at the beginning of the run and also during the run itself to double-check a possible shift of the factory settings. Standard solutions containing 0.001~mg/l, 0.005~mg/l, 0.01~mg/l, 0.10~mg/l, 1.0~mg/l, 10.0~mg/l, 20.0~mg/l, 50.0~mg/l concentration of $\rm{F^-}$ are used for absolute calibration. Fig.~\ref{fmeno_calib} shows the calibration curves. \begin{figure}[h] \begin{center} \resizebox{11cm}{!}{\includegraphics{fmeno_calib1.eps}} \resizebox{11cm}{!}{\includegraphics{fmeno_calib2.eps}} \caption{Calibration curves of fluorine monitoring sensors.} \label{fmeno_calib} \end{center} \end{figure} \newpage \section{Results and discussion} To describe the operating conditions (Tab.~\ref{summary_data_taking}) the data-taking period is divided into two runs over three years. \par \begin{table}[h] \caption{Summary table of Closed Loop (CL) and Open Loop (OL) channels.} \label{summary_data_taking} \begin{center} \begin{tabular}{|c|c|c|ccc|} \hline Run & \multicolumn{1}{c|}{Cycle} & \multicolumn{1}{c|}{Period} & \multicolumn{3}{c|}{Comment} \\ \hline 1 & 1 & 29/08/2008 - 11/10/2008 & stable currents & CL & unused filters, 9 ch CL, 2 ch OL\\ 1 & 2 & 12/10/2008 - 22/01/2009 & stable currents & CL & unused filters 9 ch CL, 2 ch OL\\ 1 & 3 & 23/01/2009 - 28/04/2009 & increasing currents & CL & unused filters 9 ch CL, 2 ch OL\\ 1 & 4 & 29/04/2009 - 14/07/2009 & increasing currents & CL & unused filters 9 ch CL, 2 ch OL\\ 1 & 5 & 15/07/2009 - 27/07/2010 & decreasing current & OL & used filters 0 ch CL, 11 ch OL\\ 2 & 1 & 28/07/2010 - 07/01/2011 & stable currents & CL & used filters 7 ch CL, 2 ch OL\\ 2 & 2 & 08/01/2011 - 05/07/2011 & increasing currents & CL & used filters 7 ch CL, 2 ch OL\\ \hline \end{tabular} \end{center} \end{table} Each run is characterized by cycles of operation. Fig.~\ref{totalcurrentplusfmeno} shows the average of all RPC anodic dark currents $I_i(t)$ over $n$ gaps, normalized by their initial values $I_i(t_0)$. The z-axis scale (color-coded) shows the F$^-$ produced by the system. The increase of dark anode currents of up-stream gaps in run 1 and run 2 is clearly visible, as well as the increase of F$^-$ concentration. The dark currents of down-stream gaps, as well as the currents of all RPCs in OL, are found stable. \par \begin{figure}[htbp] \begin{center} \resizebox{9cm}{!}{\includegraphics{totalcurrent_plus_fmeno2.eps}} \caption{Upstream gap total current during all runs with F$^{-}$ production.} \label{totalcurrentplusfmeno} \end{center} \end{figure} In run 1, the purifier cartridges are filled with unused material. Eleven double-gap RPC detectors are used, nine in CL and two in OL mode. Cycle 1 and cycle 2 have stable currents up to April 2009, when an increase in the dark current occurs for all up-stream gaps in CL, leaving the down-stream gaps stable. Cycle 4 in particular shows a clear increase of currents, and was terminated before permanent damage occurred to the detector. The lifetime of purifiers is determined by evaluating the duration of cycle 1 and cycle 2 \begin{equation} \tau_{\rm run~1} = 211 \pm 2~{\rm days} \end{equation} The total gas flow is 63$\pm$3~l/h. We measure the fluorine production (Fig.~\ref{fmeno_plot_12}) during run 1 as 1.10$\pm$0.05~$\rm{\mu mol/l}$ corresponding to a total accumulation in the CL of $(45\pm2)\,10^3~\rm{\mu mol/l}$. The purifier lifetime normalized to the $\rm{F^{-}}$ production is \begin{equation} {\hat\tau}_{\rm run~1} = 4.64\pm0.24~{\rm days/mmol/l} \end{equation} Fig.~\ref{FIG:closedloop} shows the typical behaviour of one RPC\index{RPC} detector in closed loop correlated with the concentration of the main contaminants found. \par \begin{figure}[htbp] \begin{center} \resizebox{13.0cm}{!}{\includegraphics{closedloop.eps}} \caption{Dark currents increase (run 1) in the up-stream gap and not in the down-stream gap, correlated to the detection of gas contaminants measured using the chemical analysis setup.} \label{FIG:closedloop} \end{center} \end{figure} \par Before starting run 2, all purifiers are regenerated following the CERN gas group standard procedure, i.e. by means of a flushing of hot ($215^{\circ}$C) Ar and H$_2$ mixture (80:20) for twelve hours. Nine double-gap RPC detectors are used (seven in CL and two in OL mode). During run 2 the flux is measured $\rm{\approx 54\pm3~l/h}$. The currents of down-stream gaps are found stable throughout all cycles as in run 1, while the currents of the up-stream gaps increase. As a cross-check, gas supplies of two gaps of the same RPC detector were swapped to check that in a pair of gaps only the up-stream gap showed currents increase. The lifetime of regenerated purifiers is evaluated: \begin{equation} \tau_{\rm run~2} = 160 \pm 2~{\rm days} \end{equation} \par The $\rm{F^{-}}$ production is measured 0.84$\pm$0.05~$\rm{\mu mol/l}$ (Fig.~\ref{fmeno_plot_12}), corresponding to an accumulation of 33$\pm 2)\,10^3~\rm{\mu mol/l}$. The purifier lifetime normalized to the $\rm{F^{-}}$ production is \begin{equation} {\hat\tau}_{\rm run~2} = 4.68\pm0.25~{\rm days/mmol/l} \end{equation} Analyses are in progress in order to confirm the release of contaminants observed in run 1 and shown in Fig.~\ref{FIG:closedloop}. \par The lesser $\rm{F^{-}}$ production in run 2 with respect to run 1 is interpreted as due to the smaller number of detectors used in run 2. Although the lifetime of purifiers is measured different in run 1 and run 2, the lifetime normalized to the $\rm{F^{-}}$ production is found compatible within errors, i.e., 4.64$\pm$0.24~days/mmol/l for run 1 and 4.68$\pm$0.25~days/mmol/l for run 2. \par During both run 1 and run 2 the production of $\rm{F^-}$ is efficiently depressed by the zeolite purifier as shown in Fig.~\ref{fmeno_plot_12}. The presence of an excess production of K and Ca in coincidence with the currents increase also suggests a damaging effect of HF\cite{Abbrescia:2006hk}, produced in the system, on the K- and Ca-based zeolite framework. Further analyses are ongoing to verify the presence of contaminants in run~2. \begin{figure}[h] \begin{center} \resizebox{8.3cm}{!}{\includegraphics{fmeno_primorun.eps}} \resizebox{8.3cm}{!}{\includegraphics{fmeno_secondorun.eps}} \caption{$\rm{F^{-}}$ production during run 1 (unused filters) and run 2 (used filters).} \label{fmeno_plot_12} \end{center} \end{figure} \section{Conclusions} Preliminary results on studies of contaminants, and on characterizations of materials and gas used in the CL gas system of the CMS RPC muon detector were reported. Quantitative gas chemical analysis were performed by using GC, pH sensors and contaminants detectors. The lifetime of unused purifiers is compatible with the lifetime of regenerated purifiers when normalized to the $\rm{F^-}$ produced in the system. The anodic dark current of up-stream gaps increases when purifiers are exhausted, while the down-stream gaps show stable current. This behavior is suggestive of a mechanical filtering, i.e., the first gap acts as a filter to the second gap which does not receive a polluted gas mixture. Finally, during run~1 with unused purifiers, traces of K and Ca contaminants were found in correlation with the increase of dark anodic currents. Further studies are in progress in order to ascertain the presence of such contaminants in run~2 with regenerated purifiers.
\section{Introduction} \label{sc:intro} The detection of H, O, C$^+$, and Si$^{2+}$ in the upper atmosphere of HD209458b \citep{vidalmadjar03,vidalmadjar04,linsky10}, and the tentative detection of H in the upper atmosphere of HD189733b \citep{lecavelier10} and Mg$^+$ in the upper atmosphere of WASP-12b \citep{fossati10} are among the most exciting recent discoveries related to the atmospheres of extrasolar giant planets (EGPs). The observations demonstrate that the the upper atmospheres of close-in EGPs such as HD209458b differ significantly from the thermospheres of the giant planets in the solar system. They are much hotter, they extend to several planetary radii and instead of molecular hydrogen, they are primarily composed of atoms and atomic ions. The detection of heavy atoms and ions such as O, C$^+$, Si$^{2+}$, and Mg$^+$ implies that the atmospheres of close-in EGPs are not always affected by diffusive separation. A likely explanation is that diffusive separation of the heavy atoms and ions is prevented by momentum transfer collisions with the rapidly escaping light atoms and ions. Mass fractionation during hydrodynamic escape is believed to have played an important role in the early evolution of the atmospheres of the terrestrial planets \citep[e.g.,][]{zahnle86,hunten87} but it cannot be observed in action anywhere in the solar system. Observations of EGP atmospheres thus provide a unique opportunity to study this phenomenon that should lead to a better understanding of evolutionary processes in our own solar system. Extended thermospheres give rise to large transit depths in UV transmission spectra. However, they are also potentially detectable in optical and infrared spectra. For instance, \citet{coustenis97,coustenis98} searched for an exosphere around 51 Peg b in the near-IR. In line with the current understanding, they suggested that the exospheres of close-in EGPs such as 51 Peg b are hot and composed primarily of atoms and ions. They also argued that hydrodynamic escape can lead to the escape of heavier species, and give rise to large in-transit absorption by such species in optical and near-IR spectra. However, 51 Peg b is not a transiting planet and the search for an exosphere around it was not successful. On the other hand, once the transit of HD209458b was first detected \citep{charbonneau00,henry00}, similar searches on this planet were also undertaken. \citet{moutou01} looked for visible absorption by species such as Na, H, He, CH$^+$, CO$^+$, N$_2^+$, and H$_2$O$^+$ in the upper atmosphere of HD209458b. These observations were followed by \citet{moutou03} who attempted to measure the transit depth in the He 1083 nm line that was predicted to be significant by \citet{seager00}. The most recent searches were reported by \citet{winn04} and \citet{narita05} who looked for transits in the visible Na D, Li, H$\alpha$, H$\beta$, H$\gamma$, Fe, and Ca absorption lines. So far none of the ground-based searches have led to a detection of the upper atmosphere. However, the non-detection is based on only a few observations that have proven difficult to analyze, and the search should continue. Visible and infrared observations have been able to probe the atmosphere of HD209458b at lower altitudes. In fact, HD209458b was the first EGP to have its atmosphere detected by transmission spectroscopy. The first detection was achieved by \citet{charbonneau02} who observed a deeper in-transit absorption in the Na D 589.3 nm resonance doublet compared to the adjacent wavelength bands. This detection was based on four transits observed with the Space Telescope Imaging Spectrograph (STIS) onboard the Hubble Space Telescope (HST) \citep{brown01}. The same data were later reanalyzed by \citet{sing08a,sing08b} who combined them with other observations \citep{knutson07} and created a transmission spectrum of HD209458b at wavelengths of 300--800 nm. They argued that the abundance of sodium in the atmosphere is depleted above the 3 mbar level either by condensation into Na$_2$S clouds or ionization. We note that the detection of Si$^{2+}$ in the thermosphere \citep{linsky10} constrains cloud formation mechanisms in the upper atmosphere and implies that the depletion of Na at low pressures is probably due to ionization (see Section~\ref{subsc:clouds}). The atmosphere of HD209458b has also been observed several times in the infrared with the Spitzer space telescope. \citet{deming05} detected the secondary eclipse of the planet at 24 $\mu$m by using the Multiband Imaging Photometer (MIPS). Together with similar observations of TrES-1 at 4.5 and 8.0 $\mu$m obtained by \citet{charbonneau05} who used the Infrared Array Camera (IRAC), these observations constitute the first detections of infrared emission from extrasolar planets. They were followed by \citet{richardson07} who observed the infrared emission spectrum of HD209458b between 7.5 and 13.2 $\mu$m with the Infrared Spectrograph (IRS). This spectrum was reanalyzed by \citet{swain08} who noted that it is largely featureless with some evidence for an unidentified spectral feature between 7.5 and 8.5 $\mu$m. \citet{knutson08} used IRAC to observe the secondary eclipse of HD209458b in the 3.6, 4.5, 5.8, and 8.0 $\mu$m bands. They observed a higher than expected flux in the 4.5 and 5.8 $\mu$m bands and interpreted this as evidence for the presence of a stratospheric temperature inversion that gives rise to strong water emission at these wavelengths. \citet{beaulieu10} observed the transit of the planet in the same wavelength bands and also found evidence for the presence of water vapor in the atmosphere. The detection of water vapor is also supported by \citet{swain09} who used the Near Infrared Camera and Multi-Object Spectrometer (NICMOS) on HST to observe the secondary eclipse of HD209458b between 1.5 and 2.5 $\mu$m. In addition to water vapor, the NICMOS observations revealed the presence of methane and carbon dioxide in the emission spectrum. These detections provide valuable clues to the overall composition of the atmosphere but the uncertainties in the data and degeneracies between temperature and abundances in the forward model prevent a more quantitative characterization of the density and temperature profiles. In general, difficulties associated with reducing the data and the need to describe a few uncertain data points with models of growing complexity has led to disagreements on the analysis and interpretation of transmission and secondary eclipse data on different exoplanets. The same is true of the FUV transit observations of HD209458b. \citet{vidalmadjar03} used the STIS G140M medium resolution grating to detect a 15~$\pm$~4 \% transit depth in the wings of the stellar H Lyman~$\alpha$ emission line\footnote{The core of the H Lyman~$\alpha$ line is entirely absorbed by the interstellar medium (ISM)}. Based on the data, they argued that the planet is followed by a cometary tail of escaping hydrogen atoms that are accelerated to velocities in excess of 100 km s$^{-1}$ by stellar radiation pressure \citep[see also][]{schneider98}. Later, \citet{vidalmadjar04} used the STIS G140L low resolution grating to detect absorption by H, O and C$^+$ in the upper atmosphere, and argued that the atmosphere of HD209458b escapes hydrodynamically. \citet{linsky10} used the Cosmic Origins Spectrograph (COS) on HST to confirm the detection of C$^+$ and reported on the detection of Si$^{2+}$ around the planet. They also argued that the atmosphere escapes hydrodynamically. \citet{benjaffel07,benjaffel08} disagreed with the interpretation of the G140M observations. He reanalyzed the G140M data and argued that the 15 \% H Lyman~$\alpha$ transit depth was exaggerated because the data were partly corrupted by short-term variability of the host star and geocoronal Lyman $\alpha$ emissions. He also showed that there is no evidence for a cometary tail in the data, and that absorption by H in the extended thermosphere of the planet can explain the observations. \citet{benjaffel10} reanalyzed the G140L data and reached a similar conclusion regarding H \citep[see also][]{koskinen10}. However, \citet{benjaffel10} fitted the H Lyman~$\alpha$ observations by scaling the density profiles from the model of \citet{garciamunoz07}, and argued that suprathermal O and C$^+$ are required to fit the transit depths in the O I and C II lines. They did not explain how the suprathermal atoms form and simply chose their properties to fit the observations. \citet{holstrom08} offered yet another explanation for the H Lyman~$\alpha$ observations. They argued that hydrogen atoms cannot be sufficiently accelerated by stellar radiation pressure before they are ionized by stellar X-rays and EUV (XUV) radiation. Instead, they suggested that the observed absorption arises from a cloud of energetic neutral atoms (ENAs) that form by charge exchange between the protons in the stellar wind and the escaping hydrogen. In their model, the observed absorption reflects the velocity of the stellar wind, and the data can potentially be used to characterize the magnetosphere of the planet. \citet{ekenback10} recently updated the model to include a more realistic description of the escaping atmosphere and stellar wind properties. However, both studies ignored absorption by H in the thermosphere, which is significant, and did not attempt to explain the presence of heavier atoms and ions such as O, C$^+$, and Si$^{2+}$ in the escaping atmosphere. All of the interpretations of the UV transit data require that HD209458b has a hot and extended thermosphere. In fact, \citet{koskinen10} (hereafter K10) also showed that absorption by thermal H and O in such a thermosphere explains both the H Lyman~$\alpha$ and O I 1304 \AA~transit depths. Further, their model related the observations to a few physically motivated characteristics such as the mean temperature and composition of the upper atmosphere. They also used the observations to constrain the pressure level where H$_2$ dissociates and estimated that the H$_2$/H transition should occur at 0.1--1 $\mu$bar. However, these results are based on an empirical model that was simply designed to fit the observations. One of the aims of the current paper is to show that the results are also supported by more complex physical models. We have also attempted to establish a more comprehensive description of the thermosphere that treats it as an integral part of the whole atmosphere rather than a separate entity. In order to do so, we developed a new one-dimensional escape model for the upper atmosphere of HD209458b that includes the photochemistry of heavy atoms and ions, and a more realistic description of heating efficiencies. This model is described in detail by \citet{koskinen12a} (hereafter, Paper I). We also used results from a state-of-the-art photochemical model (Lavvas et al., \textit{in preparation}) to constrain the density profiles of the observed species in the lower atmosphere. We discuss the implications of our results in the context of different observations, and show that observations of the upper atmosphere also yield valuable constraints on the properties of the lower atmosphere. \section{Methods} \label{sc:methods} \subsection{Stellar emission lines} \label{subsc:empirical} The interpretation of the FUV transit measurements relies on accurate characterization of the stellar emission lines and, provided that the ISM is optically thick over parts of the line profile, their absorption by the ISM. The observed transit depths also depend on stellar variability. In this section we discuss the properties of the line profiles, stellar variability, and absorption by the ISM. The properties of the H Lyman~$\alpha$ and the O I 1304~\AA~triplet lines (hereafter, the O I lines) have been discussed in detail before \citep[e.g.,][]{benjaffel10,koskinen10}, and we present only a brief summary of them here. However, the COS observations of the C II 1335~\AA~multiplet (hereafter, the C II lines), and the Si III 1206.5~\AA~line (hereafter, the Si III line) \citep{linsky10} have not been modeled before, and thus we constructed emission line models for these lines based on the new data. These, and the previous models for the H Lyman~$\alpha$ and O I lines, were used to calculate the predicted transit depths in Section~\ref{sc:results}. The observed line-integrated transit depths for HD209458b are listed in Table~\ref{table:models}. \subsubsection{H Lyman~$\alpha$ and O I 1304 \AA~lines} \citet{wood05} measured the H Lyman~$\alpha$ emission line of HD209458 by using the high resolution echelle E140M grating on STIS, and used the details of the line profile to constrain the column density of H in the ISM along the line of sight (LOS) to HD209458. Following K10, we adopted the reconstructed line profile and ISM fit parameters from \citet{wood05} for the transit depth calculations in this paper. The short-term variability of the H Lyman~$\alpha$ emissions from HD209458 was estimated by \citet{benjaffel07} who used the G140M data to derive a magnitude of $\sim$8.6~$\pm$~5.6~\% for this variability -- although the large uncertainty of the observations makes it difficult to separate variability from noise. There are no observations to constrain the long-term variability or center-to-limb variations of the Lyman~$\alpha$ line on HD209458. However, typical solar characteristics provide some guidance on these properties. \citet{woods00} studied the variability of solar Lyman~$\alpha$ emissions based on satellite observations spanning four and a half solar cycles between 1947 and 1999. They found that the variability ranges between 1 and 37 \% during one period of solar rotation (27 days), and the average variability during one solar rotation was found to be 9~$\pm$~6~\%. This result agrees well with the estimated variability of the Lyman~$\alpha$ emissions from HD209458. The rotation period of HD209458 is estimated to be $\sim$10--11 days \citep{silvavalio08}. The G140M observations covered three different transits and took place within a month and a half. Each observation covered approximately 2 hours in time. Thus the data can be affected by short-term variability and it is essential that such variability be properly accounted for. For this reason, we compare our models with the results of \citet{benjaffel07,benjaffel08} and \citet{benjaffel10} who analyzed the data in the time tag mode and accounted for variability before calculating transit depths. Short-term variability in the H Lyman~$\alpha$ line is mostly related to plage activity that is modulated by solar rotation while long-term variability depends on variations in both plage and active network regions \citep{woods00}. In fact, it is misleading to imagine the transits in any of the FUV emission lines as the planet crossing a smooth, uniformly emitting disk. Rather, one has to imagine the planet crossing a relatively dark disk with scattered bright regions. For instance, during solar maximum the plages\footnote{We include enhanced network in the definition of a plage here.} cover approximately 23 \% of the solar disk while active network covers about 8.5 \%. The brightness contrasts of the plages and active network are 6.7 and 3.2, respectively, when compared to the quiet sun \citep{worden98,woods00}. This means that the transit depth in the H Lyman~$\alpha$ can vary by factors of 0.4--3 as a function of time during maximum activity. Similar variability can be expected in the other FUV emission lines. The transit depth depends mostly on the path of the planet across the stellar disk although the plage and active network coverage can also change slightly with stellar rotation during the transit. This highlights the importance of light curve analysis to work out transit depths -- limited `snapshots' during transit may be corrupted by the planet either covering or not covering an active region. However, it is also possible for the transit depth to be altered even if the variations are not immediately evident in the light curves. This complicates the analysis of the observations, and underlines the need for repeated observations at different times. We note that the plage and active network coverage is significantly smaller during solar minimum than it is during solar maximum. This means that the transit depth is likely to be more stable and closer to the true transit depth during stellar minimum. Unfortunately, the activity cycle of HD209458 has not been studied in detail, and thus we have no information about it. As we noted above, the properties of the O I lines were discussed extensively by K10 who fitted parameterized solar line profiles \citep{gladstone92} to the O I lines of HD209458 that were observed with the STIS E140M grating \citep{vidalmadjar04}. We adopted the O I line profile and ISM parameters from K10 for this study, and do not discuss these properties further here. We note that the plage and active network contrasts of the O I lines are 4.2 and 1.7, respectively \citep{woods00}. These values are slightly smaller than the corresponding values for H Lyman~$\alpha$. We note that limb darkening or brightening can also affect the observed transit depths \citep[K10,][]{schlawin10}. However, center-to-limb variations are not particularly significant in the solar H Lyman~$\alpha$ and O I lines \citep{curdt08,rousseldupre85}. \subsubsection{C II 1335~\AA~and Si III 1206.5~\AA~lines} \citet{linsky10} used the medium resolution ($R \sim$~17,500) G130M grating of the COS instrument to observe four transits of HD209458b between September 19 and October 18, 2009 at wavelengths of 1140--1450~\AA. During each HST visit, they observed the star during transit, at first quadrature, secondary eclipse and second quadrature. In order to obtain the out-of-transit reference spectra, they co-added the secondary eclipse and quadrature observations from all four visits. They also co-added the in-transit spectra from all visits to create a single spectrum. The detections of the transits in the C II 1335 \AA~and Si III 1206.5 \AA~lines were compared with observations of other lines such as Si IV 1395 \AA~in which the transit was not detected. We note that the transit in the Si III line was not detected earlier by \citet{vidalmadjar04}. However, these authors use a wider wavelength interval to calculate the line-integrated transit depth and yet report a 2$\sigma$ upper limit of 5.9 \%, implying that the 3$\sigma$ upper limit includes the transit depth observed by \citet{linsky10}. As we have seen, stellar variability can also cause changes in the perceived transit depth. Therefore we are not convinced of the reality of a non-detection in the earlier STIS G140L observations. \begin{figure} \centering \includegraphics[width=0.7\textwidth]{figure1a.pdf} \includegraphics[width=0.7\textwidth]{figure1b.pdf} \caption{(a) The C II 1334.5 \AA~line of HD209458 \citep[solid line,][]{linsky10} fitted with a Voigt profile (see Table~\ref{table:lines}) and adjusted for absorption by the ISM (dotted line). We assumed that the column density of ground state C$^+$ in the ISM is 2.23~$\times$~10$^{19}$ m$^{-2}$. The relative velocity of the ISM with respect to Earth is -6.6 km s$^{-1}$ and the effective thermal velocity along the LOS to the star is 12.3 km~s$^{-1}$ \citep{wood05}. (b) The C II 1335.7 \AA~line of HD209458 fitted with a Voigt profile. Absorption by the ISM was assumed to be negligible. The model profiles were convolved to a spectral resolution of $R =$~17,500.} \label{fig:cii} \end{figure} In order to calculate transit depths in the C II and Si III lines, we created models for the stellar emission line profiles. We note that \citet{benjaffel10} also modeled the C II lines in order to fit the transit depth in the low resolution G140L data \citep{vidalmadjar04} where the main components of the C II multiplet are unresolved. Their model line profiles were also constrained by the high resolution STIS E140M observations of \citet{vidalmadjar03} that resolve the components, although this data has low S/N and it was not used for transit observations. Here we use the higher S/N COS observations that also resolve the main components of the C II lines to constrain the line profile models. The atomic line parameters for the C II and Si III lines are listed in Table~\ref{table:lines}. The C II multiplet consists of three separate emission lines. The two lines at 1335.66 \AA~and 1335.71~\AA~(hereafter, the C II 1335.7 \AA~line) are unresolved in the COS (and most solar) observations. The core of the C II 1334.5 \AA~line is strongly absorbed by the ISM whereas the C II 1335.7 \AA~line is not similarly affected. The ground state $^2 P_{1/2,3/2}$ of C$^+$ is split into two fine structure levels. Interstellar absorption of the C II 1335.7 \AA~emissions depends on the population of the $^2 P_{3/2}$ level in the ISM. In line with a similar assumption regarding the excited states of O (K10), we assumed that this population is negligible. To estimate the emission line profile from HD209458, we fitted the solar C II 1335.7 \AA~line from the SUMER spectral atlas \citep{curdt01} with a Voigt function and adjusted the result to agree with observations \citep{linsky10}. Figure~\ref{fig:cii} shows that the model line profile agrees reasonably well with the observed line profile. The fit parameters for this and other relevant emission lines are listed in Table~\ref{table:lines}. \begin{table}[htbp] \centering \caption{Atomic line parameters$^a$} \begin{tabular}{@{} lcccccc @{}} \toprule Line & $\lambda_{0}$(\AA) & $f_0$ & $A_0$(s$^{-1}$) & $\Delta \lambda_D$(\AA)$^b$ & $\Gamma$ & F(10$^{-15}$ erg cm$^{-2}$ s$^{-1}$) \\ \midrule C II & 1334.53 & 0.1278 & 2.393~$\times$~10$^{8}$ & 0.12 & 0.266 & 1.16 \\ C II$^c$ & 1335.66 & 0.01277 & 4.773~$\times$~10$^{7}$ & N/A & N/A & N/A \\ C II & 1335.71 & 0.1149 & 2.864~$\times$~10$^{8}$ & 0.16 & 0.224 & 1.66 \\ Si III & 1206.5 & 1.669 & 2.550~$\times$~10$^{9}$ & 0.12 & 0.535 & 2.09 \\ \bottomrule \end{tabular} \caption*{\small{$^a$from \citet{morton91} \\ $^b$The line profiles are given by $p(\lambda) = [F / (\Delta \lambda_D \sqrt{\pi})] V(a,u)$ where $V$ is the IDL Voigt function with $a = \Gamma/(4 \pi \Delta \lambda_D)$ and $u = (\lambda - \lambda_0)/\Delta \lambda_D$. The flux is given at Earth distance (47 parsec). \\ $^c$The emission line merges with the C II 1335.71 \AA~line.}} \label{table:lines} \end{table} Strong absorption by the ISM makes fitting the C II 1334.5 \AA~line more complicated than fitting the C II 1335.7 \AA~line. Following \citet{benjaffel10}, we estimated that the column density of ground state C$^+$ in the ISM is 2.23~$\times$~10$^{19}$ m$^{-2}$ by scaling the column density of C$^+$ measured along the LOS to Capella \citep{wood97} to the distance of HD209458. We fitted the solar C II 1334.5 \AA~line from the SUMER spectral atlas with a Voigt profile, and used the estimated column density and the results of \citet{wood05} to calculate absorption by the ISM. We then varied the total flux within the line profile until the results agreed with the observations of \citet{linsky10}. As a result, we obtained a pre-ISM flux ratio of [C II 1334.5 \AA]/[C II 1335.7 \AA] $\sim$~0.7, which agrees well with the solar value \citep{curdt01}. The model and observed line profiles are again shown in Figure~\ref{fig:cii}. We note that the ISM is optically thick at wavelengths between 1334.54 \AA~and 1335.58 \AA, which correspond to Doppler shifts of 2.2 and 11.2 km~s$^{-1}$, respectively. However, the observed flux in this region is not zero because of spectral line broadening in the COS instrument. The C II lines are formed in the upper chromosphere and lower transition region of the solar atmosphere. Similarly with H Lyman~$\alpha$, the brightest emissions are associated with plage activity \citep[e.g.,][]{athay89}. The plage and active network contrasts for the C II lines are 5.9 and 1.5, respectively \citep{woods00}. The large spatial variability makes it difficult to characterize center-to-limb variations. However, different solar observations point to approximately 40~\% limb brightening in the 1335.7 \AA~line and probably a similar variation in the 1334.5 \AA~line, with the intensity rising steadily at $\mu >$~0.6 \citep{lites78,judge03}. This brightening effect is due to the broadening of the emission line in the limb. The Si III line arises from a transition between the $^1 S$ ground state and the $^1 P$ excited state. Again, we fitted the Si III line profile from the SUMER spectral atlas \citep{curdt01} with a Voigt profile and adjusted the resulting line profile to agree with observations of HD209458 \citep{linsky10}. Figure~\ref{fig:si3} shows the observed and model line profiles, and the fit parameters are listed in Table~\ref{table:lines}. We assumed that the abundance of Si$^{2+}$ in the ISM is negligible. This is supported by a lack of detectable absorption by the ISM in the observed line profile. We note that absorption by the ISM affects the interpretation of the measured transit depths only if parts of the line profile are entirely absorbed or if the properties of the ISM change between observations. \begin{figure} \centering \includegraphics[width=0.7\textwidth]{figure2.pdf} \caption{The Si III 1206.5 \AA~line of HD209458 \citep[solid line,][]{linsky10} fitted with a Voigt profile (dotted line, see Table~\ref{table:lines}). Absorption by the ISM was assumed to be negligible. The model profile was convolved to a spectral resolution of $R =$~17,500.} \label{fig:si3} \end{figure} The solar Si III line has not been studied to the same degree as the C II lines. However, some constraints on the variability and center-to-limb variations of the Si III line have been obtained. For instance, \citet{nicolas77} used SKYLAB observations to study the center-to-limb variations of the Si III emissions from the quiet chromosphere. They found that the line is strongly limb-brightened. The total line intensity increased by a factor of 2.4 between the disk center and $\mu =$~0.73, and then again by a factor of 3 towards the edge of the limb. This means that the total intensity in the limb is a factor of $\sim$7 higher than at the disk center, and the line profile is also significantly broader at $\mu >$~0.73 with a self-reversal that does not appear at the disk center. The Si III emissions from the Sun also exhibit strong spatial and temporal variability \citep[e.g.,][]{nicolas82}. Limb brightening makes the transit depth appear smaller when the planet is covering the stellar disk while a steeper transit is seen during ingress and egress \citep[e.g.,][]{schlawin10}. On the other hand, if the planet covers active regions on the disk, the transit depth can appear significantly deeper (K10). Spatial variability and limb brightening of the emission lines on HD209458 can potentially be studied through a careful analysis of the transit light curves. Ideally, the observations should be analyzed in the time tag mode \citep[e.g.,][]{benjaffel07} to identify variations. This type of reanalysis of the COS data is beyond the scope of this paper. Instead, we use idealized transit depths based on a uniformly emitting stellar disk in Section~\ref{sc:results} to show that the optical depth of the extended thermosphere in the FUV lines is significant and that the transit in the O I, C II, and Si III lines is comparable to the transit in the H I line. We consider this sufficient for the present purposes. \subsection{Empirical model} \label{subsc:empirical_model} K10 developed an empirical model to fit the UV transit observations of HD209458b and other extrasolar planets. They argued that the H I transits can be explained in terms of three simple parameters that describe the distribution of H in the thermosphere. These parameters are the pressure $p_0$ where H$_2$ dissociates (the bottom of the H layer), the mean temperature $\overline{T}$ within the H layer and an upper \textit{cutoff level} $r_{\infty}$ based on the ionization of H. Transits in other emission lines can be explained by fitting the abundance of the heavier absorbers with respect to H. The details of this model are discussed by K10 and are not repeated here. Basically it calculates the transit depth observed at Earth orbit by assuming that the planet and its atmosphere constitute a spherically symmetric obstacle with a density profile in hydrostatic equilibrium up to the sonic point (when a sonic point exists). Absorption by the ISM and spectral line broadening within the observing instrument (STIS or COS) are taken into account. In Section~\ref{sc:results}, we compare the transit depths calculated by the empirical model with results based on the hydrodynamic model (see below) to show how the empirical model can be used to fit the data and guide the development of more complex models of the upper atmosphere. \subsection{Hydrodynamic model} \label{subsc:hydromodel} We developed a one-dimensional, non-hydrostatic escape model for HD209458b to constrain the mean temperature and ionization in the upper atmosphere (Paper I). Results from this model demonstrate that the empirical model is physically meaningful. It solves the vertical equations of motion for an escaping atmosphere containing H, H$^+$, He, He$^+$, C, C$^+$, O, O$^+$, N, N$^+$, Si, Si$^+$, Si$^{2+}$, and electrons. The model includes photoionization, thermal ionization, and charge exchange between atoms and ions. It calculates the temperature profile based on the average solar X-ray and EUV (XUV) flux and up to date estimates of the photoelectron heating efficiencies \citep[][Paper I]{cecchi09}. The lower boundary of the model is at 1 $\mu$bar and thus the model does not include molecular chemistry. This is justified because molecules are dissociated by photochemical reaction networks near the 1 $\mu$bar level \citep[e.g.,][]{garciamunoz07,moses11}. The upper boundary of the model is typically at 16 R$_p$. We placed the upper boundary at a relatively high altitude above the region of interest in this study, which is below 5 R$_p$. However, we do not consider the results to be necessarily accurate above 5 R$_p$ (see Paper I for further details). \section{Results} \label{sc:results} \subsection{Transit depths} \label{subsc:transits} In this section we constrain the temperature and composition of the upper atmosphere of HD209458b through a combined analysis of transit observations in the FUV emission lines. We also compare results from a hydrodynamic model (see Section~\ref{subsc:hydromodel} and Paper I) with the empirical model of K10, and confirm that the latter can be used to constrain the basic properties of the density profiles in the thermosphere. \subsubsection{Neutral atoms} \label{subsc:neutrals} K10 demonstrated that absorption by hydrogen in the extended thermosphere of HD209458b explains the transits in the H Lyman~$\alpha$ line, and used the observations to constrain the mean temperature and composition of the thermosphere. They fitted both the line-integrated transit depth based on the low resolution G140L data \citep{vidalmadjar04,benjaffel10}, and the transit depths and light curve based on the medium resolution G140M data \citep{vidalmadjar03,benjaffel07,benjaffel08} (see Figures 5 and 6 of K10 for the results). The results imply that the lower boundary of the absorbing layer of H is at $p_0 =$~0.1--1 $\mu$bar, the mean temperature within the layer is $\overline{T} =$~8,000--11,000 K, and the upper boundary is at $r_{\infty} =$~2.7 $R_p$. Recent photochemical calculations imply that H$_2$ dissociates near the 1 $\mu$bar level [e.g., Paper I, \citet{moses11}], and this is also supported by an observational lower limit for the vertical column density of H \citep{france10}. Hence the mean temperature in the thermosphere of HD209458b is approximately 8,250 K (the M7 model of K10). Here we show from simple arguments that the M7 model agrees in principle with the observed H Lyman~$\alpha$ transit depth. A similar procedure can also be used to obtain crude estimates of the transit depth for other systems, provided that the properties of the line profile and the ISM are known. Later in this section we compare the results of the empirical model with results from the hydrodynamic model, and show that they are consistent. To start with, the line-integrated H Lyman $\alpha$ transit depth of 6.6~$\pm$~2.3~\% \citep{benjaffel10} is consistent with a 6 \% transit depth at 1215.2~\AA~(a Doppler shift of -120 km~s$^{-1}$ from the line core) i.e., the blue peak of the observed line profile (Figure 5 of K10). Assuming that the extended atmosphere is spherically symmetric, a 6~\% transit depth measured in the blue wing (bw) of the line profile implies an optical depth of $\tau_{\textrm{bw}} \approx$~1 at the impact parameter of 2.1~$R_p$ from the center of the planet. Figure~\ref{fig:voigt} shows the absorption cross section of H in the Lyman~$\alpha$ line at a temperature of 8,250 K. The cross section at 1215.2~\AA~is $\sigma_{\textrm{bw}} =$~2~$\times$~10$^{-23}$ m$^2$ and thus a LOS column density of $N_{\textrm{H}} =$~5~$\times$~10$^{22}$ m$^{-2}$ at 2.1 R$_p$ is required to explain the observed absorption. \citet{ekenback10} and \citet{lammer11} argued that the optical depth of H in the thermosphere of HD209458b is not significant and instead a large cloud of energetic neutral atoms (ENAs) is required to explain the H Lyman~$\alpha$ observations. In particular, \citet{lammer11} claimed that a column density of $N_{\textrm{H}} \approx$~10$^{31}$ m$^{-2}$ in the thermosphere is required for strongly visible absorption. It is easy to see that this estimate is not correct because the wings of the line profile become optically thick with column densities much smaller than this. It also disagrees with \citet{garciamunoz07} and \citet{benjaffel07,benjaffel08} who were the first to suggest that H in the thermosphere may be sufficiently abundant to explain the observations. This basic result has also been confirmed by more recent calculations by \citet{trammell11}. All of these calculations show that the optical depth of the thermosphere below 3 $R_p$ is not negligible. \begin{figure} \centering \includegraphics[width=0.7\textwidth]{figure3.pdf} \caption{The absorption cross section of H in H Lyman~$\alpha$ line at $T =$~8,250~K. Thermal and natural line broadening were modeled with a Voigt profile.} \label{fig:voigt} \end{figure} The empirical model of K10 is simplified by the use of the hydrostatic approximation. This is well justified even if the atmosphere is escaping. In general, the density profile of the escaping gas in the thermosphere can be estimated from \citep{parker64a}: \begin{equation} n(\xi) c^2 (\xi) = n_0 c_0^2 \exp \left( -\int_1^{\xi} \frac{\textrm{d}u}{u^2} \frac{W^2}{c^2} \right) \exp \left(-\int_1^{\xi} \frac{du}{c^2} v \frac{dv}{du} \right) \label{eqn:parkerden} \end{equation} where $\xi = r/r_0$, $c^2(\xi) = k T(\xi)/m$, $W = G M_p / r_0$, $m$ is the molecular weight, and $v$ is the vertical flow speed. For convenience, we retained Parker's original notation. The first integral on the right hand side of equation~(\ref{eqn:parkerden}) applies in hydrostatic equilibrium. The second integral is negligible below the sonic point for transonic escape and always negligible for subsonic escape (i.e., evaporation). The sonic point on HD209458b is always above 3 $R_p$ (Paper I) and thus the density profile of H is approximately hydrostatic at least up to 2.1 $R_p$. Assuming a hydrostatic atmosphere, the LOS column density of H at 2.1 $R_p$ can be estimated from: \begin{equation} N_{\textrm{H}} (r = 2.1 \ \textrm{R}_p) \approx n_0 \exp \left[ \frac{G M_p m}{k \overline{T}} \left( \frac{1}{r} - \frac{1}{r_0} \right) \right] \sqrt{2 \pi r H( r )} \label{eqn:hcolumn} \end{equation} where $\overline{T}$ and $H$ are the mean temperature and scale height, respectively. We note that the mean thermal escape parameter is: \begin{equation} \overline{X}( r ) = \frac{G M_p m}{k \overline{T} r}. \end{equation} Assuming that $\overline{T} =$~8,250 K and using the planetary parameters of HD209458b ($M_p =$~0.7 $M_J$, $R_p =$~1.3 $R_J$), we obtain values of $\overline{X}_0 =$~13.8 and $\overline{X} (2.1 \ \textrm{R}_p) =$ 6.6. Thus, according to equation~(\ref{eqn:hcolumn}), the column density of $N_{\textrm{H}} =$~5~$\times$10$^{22}$ m$^{-2}$ that is required to explain the observations implies that $n_0 =$~3.5~$\times$~10$^{17}$ m$^{-3}$. This in turn means that $p_0 =$~0.4 $\mu$bar i.e., close to the lower boundary of the M7 model (here the agreement with the M7 model is obviously not exact because the transit depths in K10 are based on a complete forward model of the observed transit within the whole line profile). The parameters of the empirical model can be compared with corresponding values derived from the hydrodynamic models presented in Paper I (see Section~\ref{subsc:hydromodel} here for a brief summary). For instance, the mean temperature of the empirical model corresponds roughly to the pressure-averaged temperature of the thermosphere, which is given by \citep[e.g.,][]{holton04}: \begin{equation} \overline{T_p} = \frac{\int_{p_1}^{p_2} T( p ) \ \textrm{d} (\ln p)}{\ln (p_2/p_1)}. \label{eqn:meantemp} \end{equation} The hydrodynamic calculations show that the pressure averaged (mean) temperature below 3 $R_p$ based on the average solar flux varies between 6,000 K and 8,000 K for net heating efficiencies $\eta_{\text{net}}$ between 0.1 and 1. This temperature is relatively insensitive to different assumptions about heating efficiencies or the upper boundary conditions. In the reference C2 model of Paper I the mean temperature is 7,200 K. This model is based on our best estimate of the heating efficiencies that are appropriate in the strongly ionized upper atmosphere of HD209458b. With a cutoff level at 2.7 $R_p$, we obtained a line-integrated H Lyman~$\alpha$ transit depth of 4.7 \% based on the density of H in the C2 model. This value agrees with the observations to within 1$\sigma$ \citep{vidalmadjar04,benjaffel10}, but it is smaller than the transit depth of 6.6 \% predicted by the M7 model. One reason for this is the lower mean temperature of the C2 model. In order to facilitate a direct comparison between the hydrodynamic model and the empirical model, we calculated the empirical transit depth based on the mean temperature of 7,200 K (hereafter, the M7b model). The line-integrated transit depth based on this model is 5.8 \%, which is still higher than the transit depth based on the C2 model. Figure~\ref{fig:neutrals} shows the density profiles of H, H$^+$, O, and O$^+$ from the C2 model, and the density profile of H from the M7b model. The difference between the transit depths based on the empirical and hydrodynamic models arises because the C2 model has large temperature gradients (Paper I) and a gradual H/H$^+$ transition rather than a sharp cutoff. The difference does not arise because the density profile of the C2 model deviates from hydrostatic equilibrium. Given the temperature gradient in the model, the density profile is almost exactly in hydrostatic equilibrium below 3 $R_p$. In fact, the neutral density profile of the C2 model is better represented by a mean temperature of 6,300 K (not shown). This implies that the correspondence of the mean temperature of the empirical model and the pressure averaged temperature of the hydrodynamic model is relatively good but not exact. \begin{figure} \centering \includegraphics[width=0.7\textwidth]{figure4.pdf} \caption{Density profiles O and H based on the C2 model (Paper I) and the density profile of H based on the empirical model of K10 with a mean temperature of 7,200 K.} \label{fig:neutrals} \end{figure} A better agreement between the transit depths based on the C2 and M7b models is obtained with the cutoff level of the C2 model at 5 $R_p$ (see Table~\ref{table:models}). Figure~\ref{fig:cutoff} shows the line-integrated transit depth within the H Lyman~$\alpha$ line profile as a function of the cutoff level for the C2 model. This figure indicates that the transit depth increases less steeply with altitude above 4 $R_p$ than it does below this cutoff level, and saturates near 5 $R_p$. This is a natural consequence of the fact that the LOS column density decreases approximately exponentially with altitude in the lower thermosphere and the wings of the line profile become optically thin at high altitudes. \begin{figure} \centering \includegraphics[width=0.7\textwidth]{figure5.pdf} \caption{Line-integrated transit depth within the H Lyman~$\alpha$ line as a function of the obstacle cutoff level based on the C2 hydrodynamic model (Paper I). The K10 density profile applies to the M7b model (see text).} \label{fig:cutoff} \end{figure} We note that K10 already compared the results based on the M7 model with in-transit transmission as a function of wavelength within the H Lyman~$\alpha$ line profile and the light curve derived from the G140M data \citep{benjaffel07,benjaffel08}. It is not necessary to repeat a similar comparison here because spherically symmetric models that predict line-integrated transit depths that agree with the measured values are generally compatible with both the G140L and G140M data. This is partly because the uncertainty of the individual data points within the line profile and the light curve is large, and thus they do not strongly constrain the properties of the atmosphere. According to Figure~\ref{fig:neutrals}, the H/H$^+$ transition in the C2 model occurs near 3.1 $R_p$. The exact location depends on photochemistry and vertical velocity, and generally the transition occurs near or above 3 $R_p$ (Paper I). With a fixed pressure at the lower boundary, a faster velocity leads to a transition at a higher altitude. These results disagree with \citet{yelle04} and \citet{murrayclay09} who predicted a lower transition altitude, but they agree qualitatively with the solar composition model of \citet{garciamunoz07}. The density profiles of O and O$^+$ are strongly coupled to H and H$^+$ by charge exchange. As a result, the O/O$^+$ transition occurs generally near the H/H$^+$ transition. For instance, in the C2 model it is located near 3.4 $R_p$. We note that significant ionization of H and O above 3 $R_p$ is anticipated by K10 and there is good agreement on this between the empirical and hydrodynamic models. The detection of O at high altitudes constrains the mass loss rate and the ionization state of the upper atmosphere (see Section~\ref{subsc:ionescape}). However, the large uncertainty in the observations means that repeated observations are required to confirm the transit depth. The M7 model of K10 with a solar O/H ratio \citep{lodders03} yields an O I transit depth of 3.9 \%, which is within 1.5$\sigma$ of the observed value and therefore a satisfactory fit to the data. This value agrees well with the O I transit depth based on the C2 model (see Table~\ref{table:models}). K10 argued that a higher transit depth is possible if the mean temperature is higher and/or if the O/H ratio is enhanced with respect to solar. The hydrodynamic calculations indicate that the latter option is more favorable because higher temperatures lead to stronger ionization of O and may not help to significantly enhance the transit depth. Indeed, the predicted transit depth agrees with the observations to better than 1$\sigma$ if the O/H ratio is enhanced by a factor of 5 relative to solar (see the MSOL2 model in Table~\ref{table:models}). \begin{table}[htbp] \tiny{ \centering \caption{Model parameters and transit depths (\%)} \begin{tabular}{@{} cccccccccc @{}} \toprule & & & & H I$^{\textrm{c}}$ & O I & C II 1334.5 \AA & C II 1335.7 \AA & Si III \\ Model & $\dot{M}$ (10$^7$ kg~s$^{-1}$) & $\eta_{\text{net}}$$^{\text{a}}$ & $\overline{T}_p$$^{\text{b}}$ (K) & 6.6~$\pm$~2.3$^{\text{d}}$ & 10.5~$\pm$~4.4 & 7.6~$\pm$~2.2 & 7.9~$\pm$~1.5 & 8.2~$\pm$~1.4 \\ \midrule C1 & 5.6 & 0.56 & 7250 & 5.5 & 4.1 & 3.4 & 6.8 & 5.0 \\ C2 & 4.0 & 0.44 & 7200 & 5.7 & 4.0 & 3.2 & 6.7 & 4.6 \\ C3 & 6.4 & 0.57 & 6450 & 5.4 & 3.7 & 3.1 & 5.5 & 3.5 \\ C4 & 4.5 & 0.46 & 7110 & 5.2 & 3.7 & 3.2 & 6.3 & 4.6 \\ C5 & 5.6 & 0.56 & 7290 & 5.6 & 4.2 & 3.5 & 6.9 & 5.1 \\ C6 & 3.9 & 0.45 & 7310 & 5.9 & 4.1 & 3.3 & 6.9 & 4.6 \\ SOL2 & 11.0 & 0.50 & 7390 & 5.0 & 4.2 & 4.3 & 7.8 & 6.8 \\ MSOL2 & 6.0 & 0.66 & 7370 & 5.6 & 7.1 & 5.5 & 10.5 & 8.0 \\ M7 & N/A & N/A & 8250 & 6.6 & 3.9 & 3.9 & 8.0 & 5.8 \\ \bottomrule \end{tabular} \caption*{\small{$^a$Net heating efficiency (see Section~\ref{subsc:hydromodel}) i.e., the ratio of the net heating flux at all wavelengths to the unattenuated stellar flux (0.45 W~m$^{-2}$) at wavelengths shorter than 912~\AA. \\ $^b$Pressure averaged temperature below 3 $R_p$. \\ $^c$Line-integrated transit depth. The upper boundary of the absorbing atmosphere is at 5 R$_p$ apart from the M7 model where it is at 2.7 $R_p$ for neutrals and at 5 $R_p$ for ions. \\ $^d$Observed values from \citet{benjaffel10} and \citet{linsky10} with 1$\sigma$ errors.}} \label{table:models}} \end{table} We have now verified that the empirical model can be used to constrain the mean temperature and extent of the absorbing layer in the thermosphere of HD209458b. In particular, the comparison of the empirical model with the hydrodynamic model shows that the results of K10 were not affected by the simplifying assumption of hydrostatic equilibrium. We note that the purpose of the empirical model is to identify physical processes that might otherwise be missed in more complex models that are often based on a large number of uncertain assumptions. The results from \textit{any} model can now be compared with the observations by identifying the limits of the absorbing layer and calculating the global pressure averaged temperature within that layer. The values can then be compared with the parameters of the best-fit empirical model. \subsubsection{Ions} \label{subsc:ions} Figure~\ref{fig:ions} shows the density profiles of the carbon and silicon ions in the thermosphere of HD209458b based on the C2 hydrodynamic model (Paper I). The major carbon and oxygen-bearing species in the lower atmosphere, CO and H$_2$O, dissociate near the 1 $\mu$bar level \citep[e.g.,][or Lavvas et al., \textit{in preparation}]{moses11}. Thermochemical calculations indicate that SiO is the dominant silicon-bearing gas on HD209458b \citep{visscher10}. The detection of Si$^{2+}$ in the upper atmosphere implies that the formation of silicon clouds in the lower atmosphere is suppressed (see Section~\ref{subsc:clouds}), and SiO is also dissociated near 1 $\mu$bar either thermally or by photochemistry. Thus we assumed that only atomic carbon and silicon are present in the thermosphere, initially with solar abundances \citep{lodders03}. \begin{figure} \centering \includegraphics[width=0.7\textwidth]{figure6.pdf} \caption{Density profiles of C and Si in the upper atmosphere of HD209458b based on the C2 model (Paper I).} \label{fig:ions} \end{figure} According to Figure~\ref{fig:ions}, the C/C$^+$ transition occurs at a much lower altitude of 1.2 $R_p$ than the H/H$^+$ and O/O$^+$ transitions. Silicon is also almost fully ionized at the lower boundary of the model, and the Si$^{2+}$/Si$^+$ ratio is about 10 \% below 3 $R_p$. These results are in qualitative agreement with the observations, because they show that H and O are mostly neutral below 3 $R_p$ whereas C and Si are mostly ionized, and a significant abundance of Si$^{2+}$ is possible. However, it is also useful to explore if the results from the models are in quantitative agreement with the observations and if not, to adjust the model parameters as necessary to explain the data. With a cutoff level at 2.7 $R_p$, we obtained line-integrated transit depths of 2.3~\%, 3.6~\%, and 2.5~\% in the C II 1334.5 \AA, C II 1335.7 \AA~and Si III lines, respectively, based on the C2 model. Here we assumed that the population of the $^2 P$ levels of C$^+$ are in LTE. We also calculated empirical transit depths based on the M7 model. In order to do this we assumed that both C and Si are ionized at the lower boundary, and that 10 \% of silicon is Si$^{2+}$. With these assumptions we obtained C II 1334.5 \AA~and 1335.7 \AA~transit depths of 2.7~\% and 4.2~\%, respectively, and a Si III transit depth of 3 \%. These values agree well with the C2 model, and further demonstrate the consistency of the empirical and hydrodynamic models. However, they deviate from the observed values by more than 2$\sigma$. The cutoff level of the empirical model is somewhat arbitrary. For neutral species it is partly based on ionization (K10), but this criterion obviously does not apply to ions. With a cutoff level at 5 $R_p$ for the ions only, the M7 model yields line-integrated transit depths of 3.9 \%, 8 \%, and 5.8 \% in the C II 1334.5 \AA, C II 1335.7~\AA, and Si III lines, respectively, if 40 \% of silicon is Si$^{2+}$. These values agree with the observed values to better than 2$\sigma$. Similarly, by extending the cutoff level of the C2 model to 5 $R_p$, we obtained transit depths of 3.2 \%, 6.7 \%, and 4.6 \% in the C II 1334.5 \AA, C II 1335.7~\AA, and Si III lines, respectively (see Table~\ref{table:models}). These values deviate from the observed values by 2$\sigma$, 0.9$\sigma$, and 2.6$\sigma$, respectively. The transit depths predicted by the M7 model are higher partly because the mean temperature of 8,250 K is higher than the corresponding temperature in the C2 model (Paper I). This also leads to the higher Si$^{2+}$/Si$^+$ ratio that we used here. It is not clear if the apparent disagreement between the models and some of the observations needs to be taken seriously. Stellar activity and other uncertainties mean that the true transit depth can differ from measured values by a significant factor (see Section~\ref{subsc:empirical}). Further, the C2 model agrees with the line-integrated H Lyman~$\alpha$ and C II 1335.7~\AA~transit depths to within 1$\sigma$, and with the O I and C II 1334.5~\AA~lines to within 2$\sigma$. Thus we could argue that the present observations are roughly consistent with solar abundances and heating based on the average solar XUV flux. Nevertheless, we explore the apparent disagreement between the C2 model and the observations further below. This disagreement is limited to the O I, C II 1334.5~\AA, and Si III lines. Figure~\ref{fig:transits} shows the observed in-transit flux differences in the C II and Si III lines as a function of wavelength together with different model predictions. The observations indicate that the transit depths based on the C2 model fall short of the observed values because the model underestimates the width of the absorption lines. \citet{linsky10} argued that there is velocity structure within the line profiles near Doppler shifts of -10 km~s$^{-1}$ and 15 km~s$^{-1}$ that accounts for broad absorption. However, the uncertainty of the individual data points is too large to constrain the shape of the absorption lines in detail. We agree with \citet{linsky10} that the presence of velocity structure needs to be confirmed by future observations. Thus the additional absorption could also arise from spectral line broadening. \begin{figure} \centering \includegraphics[width=0.7\textwidth]{figure7a.pdf} \includegraphics[width=0.7\textwidth]{figure7b.pdf} \includegraphics[width=0.7\textwidth]{figure7c.pdf} \caption{Flux differences between the stellar in-transit and out-of-transit C II and Si III emission lines. The data points were taken from \citet{linsky10}. Also shown are model predictions based on the C2 model (dashed line) and the SOL2 model (dash-dotted line). See text and Table~\ref{table:models} for the details of these models.} \label{fig:transits} \end{figure} We note that H$_2$ absorbs within the Si III line and it can contribute to the observed absorption. The cross section has a lot of structure in this wavelength region, but for the sake of the argument we adopted a mean cross section of $\sim$2~$\times$~10$^{-23}$ m$^{-2}$ \citep{backx76}. A transit depth of 8 \% across the line profile implies that the atmosphere is optically thick up to 2.4 $R_p$ where the LOS optical thickness is $\tau \sim$~1, and thus a H$_2$ column density of 5~$\times$~10$^{22}$ m$^{-2}$ along this LOS would be required to explain the observed transit depth. This is unrealistic because the required column density is higher than the corresponding column density of H (see Section~\ref{subsc:neutrals}). Also, according to our photochemical calculations H$_2$ dissociates in the thermosphere, and the mixing ratio of H$_2$ falls below 0.1 above the 0.1 $\mu$bar level. Such a low abundance of H$_2$ has no effect on the Si III transit depth. A higher mean temperature leads to higher transit depths. Therefore we generated a new hydrodynamic model by multiplying the average solar flux by a factor of 2 and assumed a net heating efficiency of $\eta_{\text{net}} =$~0.5 (hereafter, the SOL2 model). This model agrees to better than 1$\sigma$ with all of the observed transit depths apart from the O I and C II 1334.5~\AA~lines (see Figure~\ref{fig:transits} and Table~\ref{table:models}). The mean temperature of the model below 3 $R_p$ is 7,400 K, which is lower than the mean temperature in the M7 model. However, absorption by the SOL2 model is strengthened by velocity dispersion within the escaping plasma that is not included in the M7 model. In general, the outflow velocity of the SOL2 model is significantly higher than the velocity in the C2 model. The model also predicts a mass loss rate of 10$^8$ kg~s$^{-1}$, which is twice as high as the mass loss rate based on the C2 model. This proves that a higher stellar XUV flux or a corresponding alternative energy source can explain the observations. An enhancement of the average solar flux by a factor of 2 is not unreasonable, and would roughly correspond to solar maximum conditions. In Paper I we noted that the stellar XUV flux, or the corresponding alternative heat source, would have to be 5--10 stronger than the average solar flux to produce a mean temperature between 8,000--9,000 K. Under such circumstances, the predicted transit depths in the C II and Si III lines would obviously be even higher than the values predicted by the SOL2 model. Indeed, higher temperatures broaden absorption in the wings of the line profiles and may help to explain the in-transit flux differences better (see Figure~\ref{fig:transits}). However, the energy input and temperature in the model cannot be increased without bound. Higher temperatures and flux lead to more efficient ionization of the neutral species, and as a result the transit depths in the H Lyman~$\alpha$ and O I lines begin to decrease. Also, mass loss rates of 10$^{9}$--10$^{10}$ kg~s$^{-1}$ lead to the loss of 10--100 \% of the planet's mass over the estimated lifetime of the system, and this probably limits reasonable energy inputs to less than $\sim$10 times the solar average on HD209458b. In addition to higher temperature and velocity, supersolar abundances of O, C, and Si can also lead to higher transit depths. This option is interesting because it also allows for a higher transit depth in the O I lines. As an example, we generated the MSOL2 model by enhancing the solar O/H, C/H, and Si/H abundances in the hydrodynamic model by a factor of 5. As a result, we obtained transit depths that agree with nearly all of the observed line-integrated transit depths to better than 1$\sigma$ (see Table~\ref{table:models}). However, the MSOL2 model overestimates the line-integrated C II 1335.7~\AA~transit depth, and generally overestimates absorption within the cores of the C II and Si III lines. This could imply that a higher temperature or some other source of additional broadening is a better explanation of the Si III and C II transit depths while a supersolar O/H ratio is required to match the measured O I transit depth. However, the data points in Figure~\ref{fig:transits} and the observed O I transit depth are too uncertain and do not allow for firm constraints on this. Enrichment of heavy elements is a common feature on the gas and ice giants in the solar system. For instance, the C/H, N/H, S/H, Ar/H, Kr/H, and Xe/H ratios in the atmosphere of Jupiter are all enriched by factors of 2--3 with respect to solar abundances \citep[e.g.,][]{mahaffy00,wong04}. On Saturn, on the other hand, the C/H ratio is enriched by a factor of 10 \citep{flasar05,fletcher09}. Enrichment by factors of 4--20 is expected in the N/H and S/H ratios, although condensation of NH$_3$ and H$_2$S in the deep atmosphere of Saturn makes it difficult to constrain the abundances precisely \citep[see][for a review]{fouchet09}. On Neptune and Uranus the C/H ratio is believed to be enriched by factors of 30--50 \citep[e.g.,][]{owen03,guillot07}, and similar enrichments are possible in the abundance ratios of some of the other heavy elements. Substantial enrichment of heavy elements with respect to solar abundances is therefore also feasible in EGP systems even if the metallicity of the star is close to solar. Unfortunately, we cannot use the current observations to constrain the elemental abundances of the atmosphere with accuracy. In this regard, the large uncertainty of the observations is unfortunate, because similar observations can potentially be used to estimate them. The dissociation of molecules at the relatively high pressure of 1 $\mu$bar and the lack of diffusive separation mean that the abundances of the heavy atoms and ions are simply dependent on the elemental abundances and ionization rates. Observations of the neutral species can therefore be used to constrain the temperature and ionization state, and thus the elemental abundances of the heavy species, but the S/N of the current data does not allow for strong constraints. It is interesting to note that while the velocity structure of the escaping plasma can lead to broader absorption that helps to explain the transit depths, it is not necessarily detectable in the data. For instance, Figure~\ref{fig:transits} shows the transit depths based on the SOL2 model that has a relatively high radial velocity reaching 11 km~s$^{-1}$ by 5 $R_p$. The velocity structure is not detectable because the optical depth of the high velocity material is not sufficient, the LOS velocity at the limb of the planet is slower than the radial velocities in general, and because spectral line broadening within the COS instrument smooths the structure out of the line profiles. If the presence of velocity structure is confirmed in the data \citep{linsky10}, it probably implies that there is detached, optically thick plasma moving at large velocities around the planet. If this turns out to be the case, interaction with the stellar wind probably plays a role in giving rise to the observed absorption. Such interaction may also produce turbulence that can broaden the absorption further \citep[e.g.,][]{tian05}. However, we note that non-thermal broadening such as that proposed by \citet{benjaffel10} does not appear to be necessary to explain the current observations. \subsection{Ionospheric escape} \label{subsc:ionescape} The escape of heavy atoms and ions has interesting consequences for the nature of the upper atmosphere. Here we discuss these consequences based on simple analytic arguments, and without explicit use of any complex models. The detection of heavy neutral species can be used to constrain the mass loss rate while the detection of heavy ions outside the atmosphere of the planet potentially constrains the strength of the planetary magnetic field. For instance, \citet{hunten87} derived an expression for the crossover mass limit $m_c$ for a neutral species $s$ to be dragged along by an escaping neutral species $t$ of mass $m_t < m_s$: \begin{equation} m_c = m_t + \frac{k T F_t}{n D_{st} x_t g_0 r_0^2} \label{eqn:hunten} \end{equation} where $F_t$ is the flux (s$^{-1}$ sr$^{-1}$) of species $t$, $x_t$ is the volume mixing ratio, $g_0$ is gravity at the lower boundary of the model region, and the mutual diffusion coefficient can be roughly estimated from: \begin{equation} n D_{st} = 1.52 \times 10^{18} \left( \frac{1}{M_{s}} + \frac{1}{M_t} \right)^{1/2} \sqrt{T} \ \ \ \ \textrm{cm}^{-1} \textrm{s}^{-1} \end{equation} where the masses $M$ are in amu. We used equation~(\ref{eqn:hunten}) to estimate the mass loss rate that is required to drag neutral O to the exosphere of HD209458b. Assuming that $x_t \sim$~1, $g_0 =$~10 m s$^{-1}$, $r_0 =$~R$_p$, $T =$~7,200 K, and $n D_{st} =$ 1.3~$\times$~10$^{22}$ m$^{-1}$s$^{-1}$, we obtain $F_t \approx$~2.8~$\times$~10$^{32}$ s$^{-1}$~sr$^{-1}$. This implies a minimum mass loss rate of 6~$\times$~10$^{6}$ kg~s$^{-1}$. The ionosphere of HD209458b is mostly neutral below 3 $R_p$ but even weak ionization can lead to frequent Coulomb or ion-neutral collisions that enable heavy ions or atoms to escape more efficiently. In order to illustrate the role of different collisions in transporting O and Si$^+$, Figure~\ref{fig:collisions} shows the collision frequencies for these species with H and H$^+$ as a function of altitude based on the C2 model (Paper I). We used approximate expressions for resonant and non-resonant ion-neutral collisions, and Coulomb collisions from \citet{schunk00} to calculate the momentum transfer collision frequencies. The collision frequency between two neutral species, on the other hand, was estimated from the mutual diffusion coefficient as: \begin{equation} \nu_{st} = 5.47 \times 10^{-11} \left( \frac{1}{M_s} + \frac{1}{M_t} \right)^{-1/2} \frac{n_t \sqrt{T}}{M_s} \end{equation} where the number density $n_t$ is in $cm^{-3}$. The results indicate that the transport of O depends on collisions with H below 3.5 $R_p$ whereas the transport of Si$^+$ depends on collisions with H$^+$ at all altitudes. \begin{figure} \centering \includegraphics[width=0.7\textwidth]{figure8.pdf} \caption{Momentum transfer collision frequencies based on the C2 model.} \label{fig:collisions} \end{figure} Oxygen is the heaviest neutral species detected in the escaping atmosphere, and this implies that the mass loss rate from HD209458b is $\dot{M} >$~6~$\times$ 10$^{6}$ kg~s$^{-1}$. This result agrees with \citet{vidalmadjar03} although it is less model-dependent and based on different criteria. The dominance of Coulomb collisions means that the heavy ions can escape even if the mass loss rate is lower than this. Our models predict mass loss rates of $\dot{M} \approx$~5~$\times$~10$^{7}$ kg~s$^{-1}$ (Table~\ref{table:models}) and thus diffusive separation does not take place in the thermosphere of HD209458b for neutral species with masses less than $\sim$130 amu. We note that this is the case even if escape is subsonic. In fact, equation~(\ref{eqn:hunten}) was originally derived for subsonic escape under the diffusion approximation although it is also valid for supersonic escape \citep{zahnle86}. We used the collision frequencies to derive expressions for the ion fractions $f_i = n_{\textrm{H}^+}/n_{\textrm{H}}$ at which ion-neutral and Coulomb collisions become important. The ratio of the non-resonant neutral-ion to neutral-neutral collisions exceeds 10 \% when \begin{equation} f_i \approx 10^{-24} \sqrt{ \frac{T}{M_{si} \gamma_s e^2} } \end{equation} where $i$ denotes H$^+$, $s$ denotes the colliding species, $\gamma_s$ is the neutral polarizability and all units are in cgs. However, the collision of O with H$^+$ is resonant and in this case the required ratio differs slightly from the above expression. The ratio of the Coulomb to non-resonant ion-neutral collisions, on the other hand, exceeds 10 \% when \begin{equation} f_i \approx 4.24 \times 10^{11} \frac{T^{3/2}}{Z_s^2 Z_i^2} \sqrt{ \frac{M_{sn} \gamma_n e^2}{M_{si}} }. \end{equation} where $i$ denotes $H^+$ (or the dominant ion) and $n$ is $H$ (or the dominant neutral). For Si$^+$ this fraction is $f_i \approx$~10$^{-4}$ (with $\gamma_{\textrm{H}} =$~6.7~$\times$~10$^{-25}$ cm$^3$). These equations can be used to determine if equation~(\ref{eqn:hunten}) is valid, or if more complex plasma models are required. \citet{trammell11} argued that HD209458b could have a strong planetary magnetic field that can impede the escape of ions from equatorial regions and restrict it to the polar regions. Although the magnetic field does not directly interfere with the escape of the neutral atoms, the trapped ions can stop them from escaping if the neutral-ion collision frequency is sufficiently high. Unfortunately, transit observations are not spatially resolved and they cannot be used directly to determine if escape is limited to the poles, or if the atmosphere is also escaping over the equator. However, the transit depths depend on the size of the optically thick obstacle covering the star. If mass loss is suppressed at low and mid-latitudes, the heavy species are no longer mixed into the upper atmosphere other than at the poles where they are allowed to escape. This means that the cross-sectional area covered by the ions shrinks and may become insufficient to explain the observations. Even if the plasma spreads to cover a larger area after being ejected from the poles, it is diluted in the process and thus it is not clear if the resulting cloud would have sufficient optical depth to be detectable. Assuming that the charged particles escape the Roche lobe of the planet unimpeded at the equator, we estimated an upper limit for the planetary magnetic field by evaluating the magnetic moment that allows them to do so. In order to estimate this limit, we calculated the plasma $\beta$ and the inverse of the second Cowling number (Co$^{-1}$) below 5 $R_p$ in the C2 model from: \[ \beta = \frac{2 \mu_0 p}{B^2} \ \ , \ \ \textrm{Co}^{-1} = \frac{\mu_0 \rho v^2}{B^2} \] where $v$ is the vertical velocity. Assuming a dipolar magnetic field, we obtained a limiting magnetic moment of 3.2~$\times$~10$^{25}$ Am$^2$ or 0.04 $m_J$ ($m_J =$ 1.5~$\times$10$^{27}$ Am$^2$ is the magnetic moment of Jupiter). This magnetic moment ensures that $\beta >$~10 below 5 R$_p$ and that Co$^{-1}$ reaches 10 by 5 $R_p$. Magnetic moments of $m_p \lesssim 0.04~m_J$ agree quite well with the scaling laws discussed by \citet{griesmeier04}. We note that the limiting moment produces an equatorial surface field that is approximately 4 times lower than the surface field of the Earth. \subsection{Cloud formation on HD209458b} \label{subsc:clouds} We have shown that a substantial abundance of silicon in the upper atmosphere is required to produce a detectable transit in the Si III line. If the silicon ions originate from the atmosphere of the planet, at least a solar Si/H ratio is necessary. According to thermochemical equilibrium models, silicon should condense into clouds of forsterite (Mg$_2$SiO$_4$) and enstatite (MgSiO$_3$) in the lower atmosphere of HD209458b \citep{visscher10}. If the formation of enstatite is suppressed, silicon should condense to form quartz (SiO$_2$) instead. In any case, condensation is expected to remove almost all of the silicon from the upper atmosphere. The detection of Si$^{2+}$ implies that the abundance of silicon in the thermosphere is substantial, and thus the condensation of silicon does not take place in the atmosphere of HD209458b. This has significant implications for the structure and dynamics of the atmosphere. \citet{sing08a,sing08b} analyzed the absorption line profile of Na in the atmosphere of HD209458b in detail and argued that the abundance of Na is depleted above the 1 mbar level. They suggested that this is due to the condensation of sodium into Na$_2$S, although ionization could not be ruled out decisively. Based on the condensation temperature of Na, they argued that the temperature in the upper atmosphere of HD209458b near 1 mbar is 420~$\pm$~190 K. This temperature is significantly lower than the outcome of typical radiative transfer models for HD209458b \citep[e.g.,][]{showman09}. We note that the condensation temperature of forsterite and enstatite is higher than 1,300 K at 1 mbar \citep{visscher10}. Because silicon clouds do not form, the Na$_2$S clouds cannot form either. Further, the formation of Na$_2$S relies on H$_2$S, which is dissociated above the 1 mbar level \citep{zahnle09}. This implies that any depletion of Na at high altitudes is most likely due to photoionization and/or thermal ionization. Figure~\ref{fig:condensation} shows a dayside temperature (T-P) profile for HD209458b based on \citet{showman09}. This profile is similar to the dayside temperature profile adopted by \citet{moses11}. The figure also shows the condensation curves for forsterite and enstatite. The T-P profile crosses the condensation curve for forsterite below the 100 bar level. However, the temperature profile in the deep atmosphere is uncertain, and the current profile only barely crosses the condensation curve. Also, the formation of forsterite ties only a fraction of the total abundance of silicon into clouds \citep{visscher10}. However, the T-P profile crosses the condensation curves for both forsterite and enstatite in a `cold trap' near 10 mbar. To prevent this, the temperature in the cold trap would have to be $T \gtrsim$~1,600 K. This is not totally unbelievable but probably unlikely. In any case, the temperature is close enough to the condensation curves so that moderate vertical transport might be able to preserve silicon above the cold trap. \begin{figure} \centering \includegraphics[width=0.7\textwidth]{figure9.pdf} \caption{The dayside temperature-pressure (T-P) profile of HD209458b from \citet{showman09} that includes a temperature inversion at low pressures. The condensation curves for forsterite (dotted line) and enstatite (dashed line) are shown.} \label{fig:condensation} \end{figure} \citet{spiegel09} explored a range of turbulent diffusion coefficients $K_{zz}$ that would be required to prevent the condensation of TiO and VO in the atmospheres of different EGPs, including HD209458b. In fact, they assumed that condensates form in the cold trap but are then transported to higher altitudes where they evaporate. Their work ignores the detailed chemistry of condensation, and thus the results are simply based on the ratio of $K_{zz}$ to the diffusion coefficient estimated from \[ D_p \approx v_p H \] where $v_p$ is the particle settling velocity and $H$ is the pressure scale height. The formation of condensates is probably too complicated for such a simplistic treatment, but the results provide some guidance on the value of $K_{zz}$ that is required to lift the condensates from the cold trap. We calculated $D_p$ for forsterite grains with a radius of 0.1 $\mu$m and density\footnote{This is the density of the material in the particles, not the density of the particles in the atmosphere.} of 3,200 kg~m$^{-3}$ \citep{fortney03,cooper03}. The settling velocity for such particles in the cold trap is $v_p \approx$~3~$\times$~10$^{-3}$ m~s$^{-1}$ and $D_p \approx$~2~$\times$~10$^3$ m$^2$~s$^{-1}$. Given that the upper edge of the cold trap is near the 1 mbar level where $D_p$ is higher, $K_{zz} \gtrsim$~10$^5$ m$^2$~s$^{-1}$ is sufficient to prevent the settling of the cloud particles in the lower atmosphere. We note that mass loss does not help to enhance the mixing of the particles at low altitudes significantly. The vertical velocity based on the mass loss rate of 10$^7$ kg~s$^{-1}$ is only 4.8~$\times$~10$^{-7}$ m~s$^{-1}$ at 10 mbar and 8~$\times$~10$^{-3}$ m~s$^{-1}$ at 1 $\mu$bar. Estimating $K_{zz}$ on extrasolar planets is very difficult, partly because there is no agreement on exactly what physical processes this parameter describes even in much more sophisticated solar system applications. The most recent estimates for HD209458b are based on assuming that $K_{zz} \sim \overline{v} H$ where $\overline{v}$ is the rms vertical velocity from circulation models \citep[e.g.,][]{showman09}. Based on the GCMs of \citet{showman09} and an assumed density dependence with altitude, \citet{moses11} estimated that the high pressure value of $K_{zz} =$~10$^6$ m$^2$~s$^{-1}$, which implies that $K_{zz} \gtrsim$~10$^7$ m$^2$~s$^{-1}$ at $p \lesssim$~10 mbar. If such high values are realistic, turbulent mixing is almost certain to prevent the settling of the condensates and to preserve gaseous silicon in the upper atmosphere. We note that the above requirements on the value of $K_{zz}$ may in fact be overestimated because they are based on the assumption that clouds particles form in the cold trap. Cloud formation has to be studied in the context of a photochemical model that includes the chemistry of condensation. If the chemical timescale is longer than the transport timescale, the cloud droplets may not form in the first place before the gas escapes from the cold trap. Also, the temperature structure near the cold trap needs to be constrained in greater detail. The formation of condensates is a complex problem that will be studied in future work (Lavvas et al., \textit{in preparation}) in order to better constrain the required values of $K_{zz}$. For our purposes it is sufficient to note that the current estimates of $K_{zz}$ imply, in agreement with the observations, that silicon clouds do not form in the upper atmosphere of HD209458b. \section{Discussion and Conclusions} \label{sc:discussion} We have used multiple observational constraints and theoretical models to characterize the upper atmosphere of HD209458b. Contrary to many of the earlier studies, we did not treat the thermosphere independently of the rest of the atmosphere. In fact, we have shown that observations of the upper atmosphere can be used to obtain useful constraints on the characteristics of the lower atmosphere. This is important because the extended upper atmospheres of close-in EGPs produce much larger transit depths than those arising from the lower atmosphere. In this work, we concentrated mostly on the FUV transit measurements. Future work should explore the possibility of extending the range of possible observations to other wavelength regions, as well as obtaining repeated observations in the FUV lines. Theoretical models should be developed to support new observations and to clarify the interpretation of the existing data. In agreement with K10, we showed that the H Lyman $\alpha$ transit observations \citep{vidalmadjar03,vidalmadjar04,benjaffel07,benjaffel08,benjaffel10} can be fitted with a layer of H in the thermosphere that is described by three simple parameters. These are the pressure at the bottom of the H layer, the mean temperature in the thermosphere, and a cutoff level due to ionization. The most important parameters are the pressure at the lower boundary and the mean temperature. Because H is the dominant species in the thermosphere, the data can be used to estimate the temperature of the thermosphere. Choosing a lower boundary pressure of 1 $\mu$bar based on the location of the H$_2$/H dissociation front in recent photochemical models \citep{moses11} and observational constraints \citep{france10}, we measured a mean temperature of about 8,250 K in the thermosphere below 3 $R_p$. However, the uncertainty of the observations is large, and the 1$\sigma$ upper and lower limits on this temperature are approximately 6,000 K and 11,000 K, respectively\footnote{This uncertainty does not include the possible uncertainties in the other parameters of the fit.}. We used a hydrodynamic model that treats the heating of the upper atmosphere self-consistently and the average solar XUV spectrum (Paper I) to show that a mean temperature of 8,250 K in the upper atmosphere below 3 $R_p$ is higher than the maximum temperature allowed by stellar heating. Given that a net heating efficiency equal to or higher than 100 \% is unrealistic, this temperature requires either a non-radiative heat source, additional opacity, or it implies that the XUV flux of HD209458 is higher than the corresponding solar flux. Interestingly, this would also imply that the mass loss rate could be higher by a factor of 2 or more than previously anticipated (e.g., Paper I). However, the uncertainty in the H Lyman $\alpha$ observations also allows for a slightly lower temperature of 7,200 K that is typical of stellar heating based on the average solar XUV flux and our best estimate of the net heating efficiency. Therefore the temperature implied by the H Lyman~$\alpha$ observations and the mean temperature of the basic hydrodynamic models are in good agreement. We note that the temperature in the lower thermosphere near the 1 $\mu$bar region has been constrained previously by \citet{vidalmadjar11a,vidalmadjar11b} who used the Na D lines to constrain the density and temperature profiles in the atmosphere of HD209458b. Their results point to a temperature of $\sim$3,600 K that is actually higher than the temperature at the lower boundary of our hydrodynamic model and consistent with a high mean temperature at lower pressures in the thermosphere. However, we caution the reader that the temperature profile based on the Na D lines may not be accurate. This is because \citet{vidalmadjar11a} used a simple expression for the optical depth due to Na that is based on the scale height of the atmosphere [their equation (1)]. This expression is only valid if Na is uniformly mixed with H$_2$. Since the authors also argue that Na is depleted (i.e., its mixing ratio changes with altitude) above $\sim$10 mbar, its density scale height cannot be used to estimate temperatures accurately. The detection of O in the thermosphere allowed us to constrain the mass loss rate from HD209458b based on the crossover mass concept formulated by \citet{hunten87}. This is because O is transported to high altitudes in the upper atmosphere primarily by collisions with H. We found that a minimum mass loss rate of 6~$\times$~10$^6$ kg~s$^{-1}$ is required to prevent the diffusive separation of O. This is one of the most reliable constraints on the mass loss rate available at present. Our hydrodynamic calculation based on the average solar XUV flux predicts a mass loss rate close to 5~$\times$~10$^7$ kg~s$^{-1}$. This implies that species with a mass up to $\sim$130 amu are uniformly mixed in the thermosphere. Similar constraints do not apply to heavy ions. They are transported to high altitudes by Coulomb collisions with H$^+$ that are much more efficient in preventing diffusive separation compared to collisions of neutral atoms with H. In agreement with K10, we found that the presence of O with a solar abundance and a temperature based on the H Lyman $\alpha$ measurements explains the transits observed in the O I lines. Our models predict a line-integrated transit depth of approximately 4 \% that deviates from the measured transit depth by 1.5$\sigma$ and thus agrees with the uncertainty of the observations. However, the predicted transit depths fall systematically short of the measured value. K10 suggested that a higher transit depth is possible if the O/H ratio is supersolar, the temperature of the thermosphere is higher than expected, and/or the observations probe escaping gas outside the Roche lobe of the planet, and we agree with their conclusions. As we explain below, we found that the same is true for the other heavy species. In order to calculate predicted transits in the C II and Si III lines, we created model emission line profiles for HD209458 based on SUMER observations of the Sun \citep{curdt01} adjusted to the observations of HD209458 \citep{linsky10}. We did not find evidence for significant absorption by the ISM in the C II 1335.7~\AA~or the Si III line. Parts of the C II 1334.5~\AA~line, on the other hand, are optically thick in the ISM and we took this into account in our models. We note that resolved observations of the emission line profiles can be used to characterize the composition of the ISM and the activity of the host star. This information is a valuable byproduct of the FUV transit observations that typically need to be repeated several times. With solar abundances, the same models that agree with the H Lyman~$\alpha$ transit observations tend to underestimate the transit depths in the C II and Si III lines. Similarly with the O I lines, higher transit depths in the C II and Si III lines are possible if the mean temperature of the absorbers is higher than expected and/or the C/H and Si/H ratios are supersolar. With solar abundances, a 1$\sigma$ agreement with the observed line-integrated transit depths is possible if the stellar XUV flux (or the stellar flux combined with an additional heat source) is higher than or equal to 2 times the average solar XUV flux. This corresponds to typical solar maximum conditions, and it is quite interesting that a similar enhancement may be required to explain the relatively high mean temperature in the thermosphere that we estimated earlier. Alternatively, with heating based on the average solar flux, an agreement with the observations is possible with the O/H, C/H and Si/H enhanced by a factor of $\sim$3--5 relative to solar abundances \citep{lodders03}. In any case, the atmosphere is escaping with a minimum mass loss rate given above. This is evident because of both the high temperature of the thermosphere and the detection of heavy species at high altitudes. We note that the transit observations are affected by stellar variability that can lead to significant changes in the observed transit depths. Spatial variations of intensity on the stellar disk during maximum activity or limb brightening may render the transit occasionally undetectable, or enhance it by a significant factor. Generally, observations obtained during periods of minimum activity are more reliable. \citet{benjaffel07} characterized the short-term variability of HD209458 in the H Lyman $\alpha$ line, but the variability of the O I, C II, and Si III are poorly characterized. This introduces an additional element of uncertainty into the transit depths that needs to be constrained by repeated observations of the star and the transits in the FUV lines. It also means that the qualitative explanation of the present observations that is based on heating by the average solar XUV flux and solar abundances (Paper I) may be sufficient even if the predicted transit depths in the O I, C II, and Si III lines do not exactly match the current measurements. On the other hand, if higher transit depths are confirmed, they can be used to further constrain the mean temperature and abundances as we have shown. The detection of heavy ions escaping the atmosphere can potentially be used to constrain the magnetic field strength of the planet. We estimated an upper limit of 0.04 $m_J$ for the magnetic moment that allows the heavy ions to escape unimpeded at the equator. The estimated magnetic moment agrees with the scaling laws for the magnetic field strengths of tidally locked close-in EGPs by \citet{griesmeier04}. On the other hand, a strong magnetic field inhibits the flow of ions from the equator and may only allow for escape at the poles \citep[e.g.,][]{trammell11}. If the neutral-ion collision frequencies are sufficient, the trapped ions may also prevent the neutral atoms from escaping. We note that a uniform upward flux is required to mix the atmosphere, and escape at the poles may not be sufficient to create a large enough obstacle to explain the transits in the O I, C II, and Si III lines. Detailed models of the magnetosphere that include the heavy species are required to assess if this is the case or not. The detection of Si$^{2+}$ in the upper atmosphere means that silicon cannot condense to form enstatite, forsterite, or other condensates in the lower atmosphere. This is clear because at least a solar abundance of silicon in the thermosphere is required to explain the large transit depth in the Si III line. According to the calculated temperature profiles for HD209458b \citep[e.g.,][]{showman09}, condensation is expected in a cold trap near the 10 mbar level. Provided that the temperature is not much higher than expected near the cold trap, efficient mixing is required to prevent condensation. Following an argument similar to that of \citet{spiegel09}, we estimated that an eddy mixing coefficient of $K_{zz} \gtrsim$~10$^5$~m$^2$~s$^{-1}$ below 1 mbar is sufficient to prevent the condensation of forsterite and enstatite in the cold trap. We note that much higher values than 10$^5$ m$^2$~s$^{-1}$ were assumed by recent photochemical models by \citet{garciamunoz07} and \citet{moses11}. A stratospheric temperature inversion may also be necessary to suppress condensation. This is because the cold trap cannot extend to much lower pressure than 1 mbar or the required values of $K_{zz}$ become too high. Also, the lack of condensation implies that the temperature of the lower atmosphere should be relatively high. The detection of silicon in the upper atmosphere thus provides further evidence for a stratosphere on HD209458b that was first proposed by \citet{knutson08}. We note that existing radiative transfer models do not account for molecular opacity at UV wavelengths or visible absorbers potentially generated by photochemistry \citep[e.g.,][]{zahnle09}. Our results provide motivation for more detailed models of thermal structure below the thermosphere that can constrain the chemistry of the lower atmosphere. Once this is achieved, better constraints on the dynamics can be derived. We also address an old problem related to the atmosphere of HD209458b. Based on the observed transits in the Na D lines, \citet{charbonneau02} argued that Na is depleted in the atmosphere because their cloudless solar composition model predicted significantly deeper absorption in the D lines. In addition to photoionization, molecular chemistry, and low primordial abundance of Na, they suggested that the formation of high altitude clouds can explain the observed depletion. Later \citet{sing08a,sing08b} found further evidence for the depletion of Na above the 3 mbar level, and argued that condensation of Na$_2$S is the most likely explanation. The fact that silicon does not condense implies that condensation of Na$_2$S is also unlikely. The observed depletion is therefore most likely due to photoionization and/or thermal ionization. However, the density profile and ionization state of Na should be studied in the context of molecular and ion chemistry below the 0.1 $\mu$bar level to verify that this is the case.\\ \\ We thank H. Menager, M. Barthelemy, N. Lewis, and D. S. Snowden for useful discussions and correspondence, and A. Showman and N. Lewis for sharing some of their temperature profiles. We also acknowledge the "Modeling atmospheric escape" workshop at the University of Virginia and the International Space Science Institute (ISSI) workshop organized by the team "Characterizing stellar and exoplanetary environments" for interesting discussions and an opportunity to present our work. The calculations for this paper relied on the High Performance Astrophysics Simulator (HiPAS) at the University of Arizona, and the University College London Legion High Performance Computing Facility, which is part of the DiRAC Facility jointly funded by STFC and the Large Facilities Capital Fund of BIS. SOLAR2000 Professional Grade V2.28 irradiances are provided by Space Environment Technologies.
\section{Introduction} \begin{figure} \includegraphics [width=8cm] {figure1.jpg} \caption{Energy variation of single carbon atom adsorbed on various sites of single layer, 2D hexagonal boron nitride structure (h-BN) calculated in $(4\times4)$ supercell. (a) h-BN honeycomb structure on which the adsorption energies are calculated. Nitrogen and boron atoms are represented by blue and green balls, respectively. The most favorable binding site of carbon ad-atom is marked by the red star in the figure. The path of diffusion of carbon ad-atom with the minimum energy barrier of $\sim0.65eV$ is indicated by stars. (b) Complete energy landscape of C ad-atom on h-BN structure. Light blue regions show favorable sites and the energy barrier further increases as the color goes to dark blue and purple. (c) Energy variation of carbon ad-atom is shown along the path indicated by red arrows in (a). The energy difference between the most favorable site (indicated by red star) and the bridge(Br), top of boron(TB), hollow(H), top of nitrogen(TN) sites are calculated as $0.07 eV$, $0.95eV$, $1.00eV$ and $0.03eV$, respectively.} \label{fig1} \end{figure} After the synthesis of graphene,\cite{novoselov2004, geim2007} a monolayer of $sp^2$-bonded honeycomb structure of carbon atoms, interest has focused on two dimensional nanostructures having honeycomb structures. Planar hexagonal boron nitride(h-BN) is an ionic honeycomb structure consisting of alternatively bonded boron and nitrogen atoms and is an analog of graphene.\cite{paszkowicz2002, liu2011} Despite the structural similarity, h-BN differs from graphene with its wide band gap and dielectric properties.\cite{watanabe2004} Various boron nitride structures like nanosheets,\cite{pacile2008} nanotubes\cite{chopra1995} and nanowires\cite{chen2006} have already been synthesized. Also, recent studies show that h-BN can be used to improve the current voltage properties of graphene transistors by improving the mobility of electrons in graphene as compared to graphene films on silicon substrates.\cite{dean2010} These properties hold promise for novel technological applications of h-BN structures. A thorough understanding of the properties of h-BN structure and investigating its functionalization is important for future applications of this wide band gap material. Carbon, being in the same row of the periodic table with boron and nitrogen, is one of those foreign atoms which can greatly change the physical and chemical properties of h-BN. In a recent study,\cite{ataca2010} the effects of ad-atoms adsorbed on BN were investigated using first-principles calculations, and it was shown that high coverage of carbon ad-atoms can change the magnetic properties and band gap of the system, whereas ad-atoms induce localized states in the band gap at low coverage. \begin{figure} \includegraphics [width=8.5cm] {figure2.jpg} \caption{Snapshots of the molecular dynamics simulation showing the formation of a short chain comprising four carbon atoms. The snapshots correspond to the initial, 20th, 40th and 120th steps of the molecular dynamics simulation done at 500K. Note that the formation of CAC(4) takes place as the CAC(3)leaves its initial bonding position and attaches to a single carbon ad-atom at close proximity from its top. Similar growth mechanism is also seen during the formation of CACs of length $n\leq8$.} \label{fig2} \end{figure} Single carbon ad-atom and mono-atomic chains of carbon atom are interesting entities, which can functionalize h-BN. A short segment of carbon atomic chain containing $n$ C atoms, indicated as CAC($n$) hereafter, is a truly one-dimensional carbon allotrope, and has recently drawn attention due to its linear geometry, high strength, size-dependent quantum ballistic conductance and interesting electronic properties. These properties of CACs were both theoretically\cite{abdurahman2002, tongay2004, tongay2005, senger2005, senger20051, dag2005, durgun2006, durgun20061, cahangirov2010, topsakal2010} and experimentally\cite{eisler2005, jin2009, chalifoux2010} investigated. Theoretical studies revealed the Peierls distortion in CACs,\cite{abdurahman2002, tongay2005, cahangirov2010} half-metallic and spintronic properties,\cite{dag2005, durgun2006, durgun20061} size dependent quantum conductance\cite{tongay2004, senger2005} and other geometrical structures.\cite{tongay2005} Chain structures of other group IV elements and group III-V compounds were also treated.\cite{senger20051} Concomitantly, carbon atomic chains are synthesized.\cite{eisler2005, jin2009, chalifoux2010} Recently, it was also shown that CACs can be grown on graphene and modify its properties,\cite{ataca2011} which have also been justified by the images obtained earlier using high resolution TEM.\cite{zettl2008} In this paper we studied the growth of short carbon and BN chains on single layer h-BN and graphene, respectively. We first examined the adsorption of single carbon atom on h-BN by calculating its energy landscape and diffusion barrier. This is followed by the investigation of the nucleation and growth processes of CACs on h-BN. We performed both conjugate gradient calculations and molecular dynamics simulations in order to determine the stabilities and bonding properties of these CACs and show how they can grow on the plane of BN as new carbon atoms are introduced by one atom at a time at the close proximity of an existing CAC. Once these two materials, namely carbon atomic chains and monolayer h-BN, are combined fundamentally interesting properties are attained for promising future applications. Interestingly, the properties of the grown structures depend on the number of carbon atoms in the chains, such that they exhibit an even/odd disparity. In addition, we showed that CACs grown on h-BN constitute chemically active sites for Au, Li and H$_2$, which are normally weakly bonded to h-BN. For example, H$_2$ approaching to the top of the CAC(2) is dissociated and separated into two H atoms, each attached to the chain. We also showed that CACs can attach to another h-BN from its free end to serve as pillars to increase the spacing between h-BN flakes, where molecules like H$_2$ can be stored. We presented the electronic energy band structure for various lengths of chains to see their variations with the number of chain atoms. We finally studied the problem from the reverted point of view and presented the growth patterns, optimized structural and electronic structures of BN chains grown on graphene surface. Our results revealed the unique self-assembly character of carbon and BN chains on single layer honeycomb structures and interesting features attained thereof. \section{Method} In our calculations we use the state-of-the-art first-principles plane-wave calculations within the density-functional theory\cite{kohn1965,payne1992} combined with ab-initio, finite temperature molecular-dynamics calculations using projector augmented wave potentials.\cite{blochl1994,kresse1999} The exchange correlation potential is approximated by the generalized gradient approximation with Van der Waals correction.\cite{perdew1992,grimme2006} Numerical computations have been carried out by using VASP software.\cite{VASP} A plane-wave basis set with energy cutoff of $600 eV$ is used. The Brillouin zone is sampled in the \textbf{k}-space within the Monkhorst-Pack scheme,\cite{monkhorst1976} and the convergence of the total energy and magnetic moments with respect to the number of \textbf{k}-points are tested. The convergence for energy is chosen as $10^{-5}eV$ between two consecutive steps. In the relaxation of structures and band structure calculations, the smearing value for all structures is taken as $0.01eV$. We consider adsorption of chains on $(4\times4)$ supercells and treat the system using periodic boundary conditions. The pressure on each system was kept smaller than $\sim2kBar$ per unit cell in the calculations. In the ab-initio MD calculations the time step was taken as $2.5fs$ and atomic velocities were renormalized to the temperature set at $T=500K$ and $T=1000K$ at every $40$ time steps. In the MD stability tests, the simulations were run for $10ps$. \section{Adsorption of single carbon ad-atom on h-BN} Before going into detailed studies of carbon chains, we first investigate the adsorption and migration of single carbon ad-atom, which is the starting point of CAC growth on h-BN substrate. We use a $(4\times4)$ supercell of h-BN that consists of $16$ boron and $16$ nitrogen atoms. There is an energy difference of $\sim0.2eV$ between the spin polarized and spin unpolarized energy values in favor of the spin polarized state, indicating that the system has a magnetic ground state. Therefore all of the calculations mentioned hereafter are performed using the spin polarized conditions. The most favorable binding site of single carbon atom was determined by placing the ad-atom initially to various adsorption sites at a height of $\sim2$\AA~ from the BN atomic plane and running fully self-consistent geometry optimization calculations by keeping the ad-atom fixed in $x-$ and $y-$ directions and letting the vertical $z-$coordinate of the ad-atom, which is its height from the plane, free. Meanwhile, the atoms in the BN supercell are relaxed in all directions except for one corner atom of the supercell, which is fixed in all directions to prevent h-BN from sliding. In Fig.~\ref{fig1}(a) the nitrogen and boron atoms of the optimized h-BN structure are separated from each other by $1.45$\AA. The most favorable bonding site of single carbon ad-atom, which turns out to be near the top site of nitrogen atom, is marked with a red star. The migration(diffusion) path of carbon ad-atom on h-BN with a minimum energy barrier is shown by black stars. The minimum energy barrier is calculated as $0.68 eV$.\cite{canbn} The energy landscape calculated over the whole BN hexagon also shows that the energy barrier to the diffusion further increases as the carbon atom moves away from the nitrogen atom as shown in Fig.~\ref{fig1}(b). The variation of energy calculated along the 2D path shown in Fig.~\ref{fig1}(a) is presented in Fig.~\ref{fig1}(c). As indicated in the figure, the most favorable site for the carbon atom is near the top nitrogen site, although not exactly on top of nitrogen. The energy barrier between the most favorable site and the top bridge(BR), top boron(TB), hollow(H) and top nitrogen(TN) sites were calculated as $0.07eV$, $0.95eV$, $1.00eV$ and $0.03eV$, respectively. In addition to the diffusion path analysis of a single carbon ad-atom, we next study the interaction between two carbon atoms on h-BN surface. When the distance between the two ad-atoms becomes less than a threshold distance of $\sim2$\AA, these two carbon atoms attract each other and form a CAC(2) perpendicularly attached to h-BN. This is indeed the nucleation for longer CACs. The most favorable binding site CAC(2), is again near the top site of nitrogen. A complete site analysis was also performed to confirm this result, by placing a CAC(2) on various adsorption sites and comparing the total energy values. \section{Growth of carbon atomic chains on h-BN} \subsection{Chain growth and even/odd disparity} \begin{figure} \includegraphics [width=8.5cm] {figure3.jpg} \caption{Binding energies ($E_b$), and the heights($h$) of odd and even numbered CAC($n$)'s from the atomic plane of BN are shown in green, red and blue lines, respectively. The $h$ values exhibit an even/odd family behavior depending on the number of carbon atoms in the chain. The sudden peak in the binding energy arises from the change of the magnetic state of CAC(2) from magnetic to nonmagnetic when it binds to hexagonal BN.} \label{fig3} \end{figure} Growth of the CAC further continues when a third carbon ad-atom is introduced at the close proximity of CAC(2). However, this time a complete site analysis of CAC(3) shows that the most favorable bonding site is the top of boron atom, instead of the nitrogen site. The formation of CAC(3) happens as follow: CAC(2) leaves its initial bonding position, moves higher from the BN plane and in the mean time it gets closer to the single ad-atom until they are bound to each other near the new energetic site, which is the top of boron site. Similar chain growth behavior is also seen during formation of CACs at different lengths. This process is further investigated with ab-initio molecular dynamics simulations at 500K and the snapshots taken from the growth of CAC(4) is presented in Fig.~\ref{fig2}. We initiate the MD simulation by placing a carbon ad-atom and a CAC(3) to their bonding sites as shown in Fig.~\ref{fig2}(i). The simulation was run for 2000 time steps and snapshots taken from the initial, 20$^{th}$, 40$^{th}$, and 120$^{th}$ time steps are shown. As the simulation proceeds to the 2000$^{th}$ step, the chain stays in its position shown Fig.~\ref{fig2}(iv), which is an indication of its stability at that bonding site. \begin{table} \caption{Most favorable binding sites, total binding energies($E_b$), magnetic moments($\mu$), heights($h$) of CAC($n$) from the BN plane, and the distances of the lowest carbon atom of the chain from the nitrogen($d_{C-N}$) and the boron($d_{C-B}$) atom in the BN plane for different $n$ values of carbon chains. The bonding sites and magnetic properties of CACs on BN exhibit an even/odd disparity. With the exception of the single carbon ad-atom, even numbered CACs bind to BN near the top of nitrogen(TN) atom and the odd numbered CACs bind near the top of boron(TB) atom. Additionally, the even and odd numbered chains grown on BN have magnetic and nonmagnetic(NM) ground states, respectively, with the exception of CAC(1) and CAC(2) cases. All calculations were performed on a $4\times4$ supercell which contains 32 carbon atoms.} \label{table: 1} \begin{center} \begin{tabular}{ccccccc} \hline \hline Structure & Site & $E_b(eV)$ & $\mu$($\mu_B$) & $h$(\AA) & $d_{C-N}$(\AA) & $d_{C-B}$(\AA) \\ \hline BN+C & $\sim$TN & 1.36 & 2.00 & 1.50 & 1.59 & 1.73 \\ \hline BN+CAC(2) & $\sim$TN & 2.19 & NM & 1.32 & 1.46 & 1.83 \\ \hline BN+CAC(3) & $\sim$TB & 0.28 & NM & 1.59 & 2.16 & 1.62 \\ \hline BN+CAC(4) & $\sim$TN & 0.24 & 2.00 & 1.37 & 1.50 & 1.63 \\ \hline BN+CAC(5) & $\sim$TB & 0.20 & NM & 1.60 & 2.28 & 1.62 \\ \hline BN+CAC(6) & $\sim$TN & 0.18 & 1.97 & 1.41 & 1.52 & 1.58 \\ \hline BN+CAC(7) & $\sim$TB & 0.14 & NM & 1.67 & 2.41 & 1.64 \\ \hline BN+CAC(8) & $\sim$TN & 0.08 & 1.96 & 1.48 & 1.56 & 1.59 \\ \hline \hline \end{tabular} \end{center} \end{table} \begin{figure*} \includegraphics [width=16cm] {figure4.jpg} \caption{Side and top views of the most favorable binding configurations of CAC($n$) on hexagonal BN are shown in (a) and (b). N, B, and C atoms are represented by blue, green and brown balls, respectively. CAC($n$)'s having an even number of carbon atoms (even $n$) is bound to BN near the top of nitrogen atom, whereas CAC($n$)'s with odd number of carbon atoms (odd $n$) prefer top boron site, with the exception of single carbon ad-atom. The geometries are calculated for a $(4\times4)$ supercell and their stabilities are also tested with MD simulations at $T=500K$ for $10ps$. In (b), only the carbon atom of the chain that is closest to the BN plane is shown.} \label{fig4} \end{figure*} We further continue the above analysis with fourth, fifth, sixth and seventh carbon ad-atoms. In general, once a new carbon atom is implemented at the close proximity of the existing chain CAC($n-1$), this chain leaves its binding site and is bound to the top of new carbon ad-atom in their new most favorable binding site. Namely, this binding site keeps changing between the near top nitrogen site and the near top boron site. With the exception of the single carbon ad-atom, we observe an even/odd disparity of the binding site depending on the number of atoms in the CAC, that is even numbered CACs bind to h-BN substrate near the top nitrogen site, and the odd numbered CACs bind near the top boron site. This situation can be attributed to the charge density of CAC(n) at its end, which depends whether $n$ is even or odd. The binding energies of CACs attached to graphene are calculated using the expression $E_{b}=E_{T}$[free-linear CAC(n)]+$E_{T}$[bare h-BN]-$E_{T}$[CAC(n)+h-BN]. In this expression, $E_{T}$[free-linear CAC(n)] is the total energy of fully relaxed linear chain of length $n$, $E_{T}$[bare h-BN] is the total energy of bare $(4x4)$ supercell of h-BN and $E_{T}$[CAC(n)+h-BN] is total energy obtained when CAC(n) binds to the supercell of h-BN. The complete list of these binding energies and the binding geometries of these CACs are presented in Table \ref{table: 1}. It is seen that, with the exceptional case of CAC(2), which is further discussed in the following section, the binding energies($E_b$) of the CACs tend to decrease as the length of the chains (or the number of carbon atoms) increase. Similarly, due to the decrease in the binding energy, the height($h$), namely the distance between h-BN plane and the first carbon atom of the CAC, increases. However, the $h$ values also show an even/odd disparity. In other words, there is an increasing trend in the $h$ values when the CACs are grouped as even and odd numbered chains, but the height of CAC($2n$) is always less than CAC($2n-1$), as shown in Fig.~\ref{fig3}. This situation can be explained by the fact thats the C-N bond length is shorter than the C-B bond length, and CAC(2n) is a more stable structure since it has a more symmetrical charge distribution as compared to CAC($2n-1$). These decrease $h$ distances of the even numbered CACs that are bound to the nitrogen atom. We finally test the stabilities of these h-BN+CAC($n$) complexes at finite temperatures with molecular dynamics simulations. The ab-initio MD calculations were carried out for $10ps$ at $T=500K$. The final stable bonding configurations of various CACs are shown in Fig.~\ref{fig4}. After MD simulations, these CACs on h-BN remain stable and they don't detach from the BN plane, but they slightly pull the nearest B and N atoms of h-BN upwards from the plane, as seen from the side views given in Fig.~\ref{fig4}(a). Also, the chains are not perfectly perpendicular to the BN plane, but are slightly tilted at different angles. Fig.~\ref{fig4}(b) shows the top views of the CACs, where only the carbon atoms closest to the BN planes are shown. As noted above, the even/odd disparity in the bonding sites of the chains with the exception of single carbon atom can be seen. It should be noted that ab-initio MD simulations carried out for 10ps is not sufficiently long to accumulate enough statistics. However, calculations at high temperatures are worthwhile to give a hint about the stability at relatively lower temperatures. An unstable chain would easily breakdown or detach from substrate. This did not happen in this case. Additionally, calculations using conjugate gradient method resulted in chain structures bound to the substrate with a significant binding energies, especially for short CACs. Hence, these tests provide sufficient evidence that the chains we studied are stable and remain attached to the single layer h-BN. \subsection{Electronic and magnetic properties} Having found the structural properties of CACs on h-BN, we next focus on the variations of magnetic and electronic properties of h-BN with various lengths of CACs attached. First, we perform both spin polarized and spin unpolarized energy minimization calculations for free(bare) CACs without attaching them to h-BN and compare the minimum energies of magnetic and non-magnetic cases. In a similar manner with the optimized structures of CACs attached to h-BN, it turns out the magnetization of free CACs also depends on the number $n$ of carbon atoms. Namely that a free chain has magnetic ground state if it has even number of carbon atoms, and a non-magnetic ground state if it has odd number of carbon atoms with the exception of single carbon atom in vacuum. Such an even/odd disparity in the magnetic moments of finite size CACs was previously reported.\cite{cahangirov2010} Hence, when these chains are adsorbed to h-BN sheet, the overall magnetic properties of the h-BN+CAC($n$) structures are maintained with the exception of CAC(2). While all other CACs preserve their magnetic ground states when bonded to h-BN, CAC(2) flips from the magnetic ground state to the non-magnetic ground state. This change in the magnetic moment of CAC(2) increases its bond strength with h-BN and causes a higher binding energy. This is the reason for the sudden jump in Fig.~\ref{fig3}. \begin{figure} \includegraphics [width=8.5cm] {figure5.jpg} \caption{Variation of C-C bond lengths in the short carbon atomic chains CAC($n$) grown on BN. While the even numbered chains (blue) have alternating short and long bonds, the odd chains(red) have a symmetrical bond length variation around their central atom. Note that the number of bonds in a chain is $n-1$} \label{fig5} \end{figure} \begin{figure*} \includegraphics[height=10cm]{figure6.jpg} \caption{(a) Electronic energy band structures of CACs grown on h-BN calculated for $n$= 1, 2, 3 and 4. In the magnetic cases, spin up and spin down bands are represented by blue and green lines, respectively. The bands below the red dash-dotted lines are fully occupied. The localized impurity states appearing as flat bands in the band gap originate from the $p$ bands of the carbon atoms that are at the edges of chains. (b) Isosurfaces of the difference charge densities of chains where yellow and green regions designate charge accumulation and charge depletion, respectively. The isosurface values are taken as 0.01 electron/\AA$^3$ for C, C$_2$, C$_3$ and as 0.005 electron/\AA$^3$ for C$_4$.} \label{fig6} \end{figure*} The disparities observed so far are closely related to the bond length variation of CACs with even and odd number of atoms.\cite{cahangirov2010} First, we note that the final stable structures of short CACs grown on h-BN have nonuniform bond length distribution. While the even numbered chains have polyyne type of structure with alternating short and long bonds, the odd numbered chains have symmetrical bond length variations around their center points. In other words, the bond length variations for the odd chains reach to zero at the center, whereas the variation of bond lengths continues in chains having even $n$. This situation is related to the geometrical linear symmetry of the chain as also observed in a recent study for bare chains.\cite{fan2009} For the odd chains, the geometrical center of the chain is the middle carbon atom since there are even number of bonds. On the other hand, even chains have odd number of bonds and the geometrical center is on the middle bond. Since these chains are bonded to h-BN substrate only from one side they are free to move in the other direction. Hence, the chains need to be symmetrical around the center, and symmetrically equivalent bonds must have the same length. As a result, the two equivalent bonds around the center atom of the odd chains have the same length. Variations of bond lengths of h-BN+CAC($n$) for $n$=5-8 are presented in Fig.~\ref{fig5} This overall symmetry in the geometrical structure of the odd numbered chains also leads to a symmetry in the distribution of the magnetic moments, making the ground states singlet for odd chains. In the singlet state, magnetic moments exerted by the electrons cancel each other leading to zero net spin and zero magnetic moment. Therefore, these structures have a nonmagnetic ground state. On the other hand, even chains have an uneven distribution of magnetic moments on atoms leading them to have magnetic ground states. This magnetic preference also alters the binding sites of the chains accordingly. One would expect this situation to be reverted for chains saturated from both ends because the boundary conditions for the chains are altered as shown in a recent study.\cite{cahangirov2010} An individual CAC($n$) attached to h-BN can modify the electronic band structure of h-BN. If an adsorbed CAC($n$) is sufficiently far from others, it gives rise to localized states in the band gaps and resonance states in the band continua of h-BN. In this study we consider a single CAC($n$) adsorbed to h-BN using a model where one CAC($n$) is attached to each repeating ($4\times 4$) supercell with the condition that interactions between adjacent CACs are negligible. This model recovers approximately the ($4 \times 4$) folded bands of bare h-BN, except that the energy difference between the top of the valance bands and the bottom of the conduction bands gradually increases from $4.5eV$ to $4.8eV$. Additionally, flat bands due to CAC($n$) occurs in the band gap. These flat bands actually corresponds to the localized states due to CAC($n$). In Fig.~\ref{fig6}(a) we present the electronic energy structure of h-BN+CAC($n$) calculated for $n$=1,2,3 and 4 using supercell geometry explained above. The energy positions of the filled and empty flat bands in the gap are closely related with the energy levels of the corresponding CAC($n$), which vary with $n$. For magnetic $n$=1 case, spin-up states (i.e. flat bands) originating mainly from C-$p_{xy}$ orbitals occur above the top of valence band. Empty spin-down states of the same orbital character occur near mid gap. In the non-magnetic case of $n$=2 spin-degenerate C-$p_{xy}$ and C-$p_{z}$ states are filled and occur near the top of the valence band. While a filled C-$p_{xyz}$ state touches the top of valence band both for non-magnetic $n$=3 case and magnetic $n$=4 case, the positions of empty C-$p_{xy}$ states in the band gap strongly depend on $n$. For all $n$'s the resonance states occur in the valence band within the energy of 2eV from the top. The difference charge density is calculated by subtracting the charge densities of bare h-BN and bare CAC($n$) from that CAC($n$)+h-BN system by keeping the atomic configuration of adsorbed and bare CAC unaltered. In Fig.~\ref{fig6}(b) the isosurfaces of difference charge density indicate that important changes from the corresponding bare CAC is found where CAC is bound to h-BN. \subsection{Applications of CACs on BN} \begin{figure*} \includegraphics [width=18cm] {figure7.jpg} \caption{Functionalization of BN sheets through adsorption of carbon chains. For example, a CAC(2), which is strongly bound to h-BN, creates chemically active sites for Au, Li and H atoms. H$_2$ molecule approaching to CAC(2) from sides dissociated to form two C-H bonds, whereas O$_2$ remains totally inactive. Ti atom takes the carbon atoms with itself and forms TiH$_2$. } \label{fig7} \end{figure*} Apart from its interesting and fundamental aspects, the growth of CACs on single layer graphene is a reality and they can modify graphene's physical and chemical properties. Experimental observations and images obtained using high resolution TEM\cite{zettl2008} indicates that CACs can be formed on monolayer graphene surfaces. In this study, we show that BN is also a suitable substrate for the growth of CAC($n$). Previous experimental studies\cite{liu2011} and density functional results\cite{ozcelik2012, haghi2012} have shown that epitaxial graphene grown on boron nitride surface is possible by means of chemical vapor deposition technique or exposure of BN layer to carbon atoms or dimers at long time intervals. In these studies, it was noted that defected graphene structures with undesired chains can also form during the growth of graphene.\cite{jin2009, topsakal2010} Thus, since the temperature range of our MD calculations are suitable for atomic layer deposition techniques, it is conjectured that similar atomic layer deposition or molecular beam epitaxy methods can be used for chain growth on BN layers at specific ambient conditions. \begin{figure} \begin{center} \vspace{1cm} \includegraphics [width=8cm]{figure8.jpg} \end{center} \caption{CAC(2) and CAC(3) grown between two BN flakes. The optimized spacing between the flakes increase from $3.1$\AA~ to $4.34$\AA~ and $5.82$\AA~ upon the formation of chains.} \label{fig8} \end{figure} BN is in general not a chemically active material, however with the inclusion of certain ad-atoms its chemical activity can be improved. CACs grown on BN can be used for functionalization of h-BN layers which will enable the adsorption of certain ad-atoms like hydrogen, lithium and gold. This functionalization will especially be useful where the contact on an insulating surface like BN is desired along with connection of electrodes. In order to increase the chemical activity of the inactive BN layer we use CAC(2) chain. In the previous sections, we showed that CAC(2) binds to h-BN with a very high binding energy as compared to other CACs. Besides being bonded strongly to h-BN, it also creates a chemically active sites for the absorption of other ad-atoms. Here we test the adsorption of Au, Li, Ti, H$_2$ and O$_2$ on CAC(2) and show that the chemical activity of BN can be enhanced through CAC(2) attached to it. Ad-atoms were placed in the vicinity of CAC(2) on h-BN and fully self consistent calculations were performed. Before introducing the ad-atoms, the CAC(2) was placed freely in its most favorable position as calculated in the previous section. It was seen that Au and Li atoms move towards the chain and are bound to the chemically active site of CAC(2) with binding energies of $1.77$ and $1.53eV$, respectively. These are significantly higher than the binding energies of Au and Li on bare h-BN, which are calculated as $0.01$ and $0.12eV$, respectively. This is an important result indicating that gold electrodes can easily be connected to semiconducting h-BN through CACs. A hydrogen molecule approaching to CAC(2) from the sides moves upwards and dissociates to form two C-H bonds. The binding energy for H atoms is calculated as $3.62eV$. However, when O$_2$ is introduced to the system instead of H$_2$, it stays completely inactive and stay away from CAC(2). Finally, we consider Ti ad-atom. Normally, Ti is bound to bare h-BN with an energy of $0.80eV$. However, when h-BN is functionalized with CAC(2), although Ti ad-atom initially is bound to the carbon atoms, it doesn't stay there but takes away these CAC(s) from the BN layer and forms a TiC$_2$ structure which moves away from BN. This final configuration is energetically more favorable than Ti binding to h-BN by $5.9eV$. The final relaxed geometries of all these structures are presented in Fig.~\ref{fig7}. Pillared graphene structures and CACs passivated by graphene surfaces from both ends have also been subjects of recent interest of both theoretical and experimental studies.\cite{dimitrakakis2008} It was recently shown that CACs passivated by graphene flakes from both ends produce highly stable chain structures.\cite{ravagnan2009} Oxidized graphene pillars were also shown to be useful materials for storage applications.\cite{burress2010} Motivated by previous studies, we also calculated the bonding geometries of CACs between BN layers and showed that CACs can also be grown between two BN flakes as shown in Fig.~\ref{fig8}. We have demonstrated this situation by calculating optimized bonding configurations and the spacing between two BN flakes when CAC(2) and CAC(3) are grown between them. CAC(2) is bonded to nitrogen atom from one side and to boron atom on the other side. On the other hand, CAC(3) is bonded to both of the flakes from the top of boron atom. Once CAC(2) and CAC(3) are grown, the spacing between the flakes increases from $3.1$\AA~ to $4.34$\AA~ and $5.82$\AA, respectively. The formation energies, calculated by subtracting the energy of two BN planes and the energy of CAC from the energy of the final structure, are found as $2.0$ and $0.32eV$ for CAC(2) and CAC(3), respectively. \begin{table} \caption{Most favorable binding sites, binding energies($E_b$) and the heights($h$) of B, N ad-atoms and BN or NB atomic chains from the graphene plane. Here, for example, BN(3) indicates the BN chain consisting of three atoms (namely N-B-N) grown on graphene with N being attached to graphene.} \label{table: 2} \begin{center} \begin{tabular}{cccc} \hline \hline Structure & Binding site & $E_b(eV)$ & $h(\AA)$ \\ \hline B-ad-atom & Bridge & 1.24 & 1.45 \\ \hline N-ad-atom & Bridge & 1.05 & 1.25 \\ \hline NB(2) & Bridge & 2.58 & 1.46 \\ \hline BN(2) & Top & 2.42 & 1.32 \\ \hline NB(3) & Bridge & 1.18 & 1.52 \\ \hline BN(3) & Top & 0.90 & 1.31 \\ \hline NB(4) & Bridge & 0.82 & 1.58 \\ \hline BN(4) & Top & 0.46 & 1.39 \\ \hline NB(5)& Bridge & 0.73 & 1.63 \\ \hline BN(5) & Top & 0.21 & 1.48 \\ \hline NB(6) & Bridge & 0.51 & 1.67 \\ \hline BN(6) & Top & 0.15 & 1.57 \\ \hline \hline \end{tabular} \end{center} \end{table} \section{BN atomic chains grown on graphene} Having shown the stable carbon chain structures grown on BN, we next investigate the growth of short BN chains on graphene to see if a similar self-assembly mechanism is also present for this reverted situation. Here we adopt the representation where BN($n$)+graphene indicates a BN chain consisting of $n$ atoms that is attached to graphene through N atom. For example, NB(3) + graphene (or NB(3) only) stands for B-N-B chain attached to graphene through B atom. Following the similar procedures described in Sections III and IV, we calculate most favorable binding sites, electronic properties, binding energies. We first perform MD simulations to see the self assembly mechanisms of BN chains. To this end we start by placing single B and N ad-atoms on graphene supercell and run MD simulations at 500K for 10ps. Eventually, the two ad-atoms form a BN molecule perpendicularly attached to graphene. When a third ad-atom N is placed on the graphene at the close proximity of existing BN molecule, MD simulations show that a chain comprising three atoms, B-N-B forms. This procedure was repeated for different lengths and it was observed that B and N atoms are indeed self-assembled on graphene honeycomb structure to form short BN atomic chains. Snapshots taken from these MD simulations are presented in Fig.~\ref{fig9}. \begin{figure} \begin{center} \includegraphics [width=8cm]{figure9.jpg} \end{center} \caption{Growth dynamics of short NB chains on graphene. Once a new ad-atom is placed at a random point in the neighborhood of a NB chain having $n$ atoms, the structure rearranges itself to form a chain comprising $n+1$ atoms. The snapshots are taken from the quantum MD simulations performed at 500K for 10ps. Blue, green and brown balls are, respectively, nitrogen, boron and carbon atoms.} \label{fig9} \end{figure} \begin{figure*} \includegraphics [width=16cm]{figure10.jpg} \caption{The most favorable binding configurations of short NB and BN chains grown on graphene. A NB($n$) chain is attached to graphene at the bridge site, since its nearest atom to graphene is B, and BN($n$) is bound at the top of carbon atom, since its nearest atom to graphene is N. Energies are calculated using a (4x4) supercell and their stabilities are tested with MD simulations at T = $500K$ for $10$ ps. Note that the most favorable binding site is the bridge site for both single N and single B ad-atoms. Blue, green and brown balls are, respectively, nitrogen, boron and carbon atoms.} \label{fig10} \end{figure*} \begin{figure} \begin{center} \includegraphics [width=8.5cm]{figure11.jpg} \end{center} \caption{Electronic energy structures: (a) Single B ad-atom adsorbed on graphene. (b) Single N ad-atom adsorbed on graphene. c) NB chain consisting of two atoms grown on graphene with B atom being closest to graphene. d) BN chain of two atoms grown on graphene with N being the closest ad-atom to graphene. The zero of energy is set to the Fermi energy, shown by red dash-dotted line. In the magnetic cases, spin up and spin down bands are represented by blue(dark) and green(light) lines, respectively.} \label{fig11} \end{figure} The most favorable binding sites of short BN chains on graphene and their corresponding binding energies were calculated with self consistent conjugate gradient calculations. It was found that, the binding site of a BN chain on the graphene depends on the type of its closest atom to the graphene. A NB($n$) chain is bound to the graphene layer at the bridge site, and a BN($n$) is bound at the top site as shown in Fig.~\ref{fig10}. However, both single B and single N ad-atoms are bound at the bridge site. Contrary to the CAC case, no clear even/odd disparity was obtained in h-BN+graphene complexes owing to the ionic bonding between B-N atoms. Here there is another effect related with the number of electrons in the chain. Chains having odd number of atoms have odd number of electrons, while even numbered chains have even number of electrons. The binding energies of these chains shown in in Table \ref{table: 2} suggest that BN chains are bound with higher energies to the graphene as compared to the binding energies of CACs on BN, especially for longer chains. The binding energy of a boron ending chain (i.e. NB) is always higher than a nitrogen ending chain (i.e. BN) site. Here we note that the supercell size and adsorbate-adsorbate The even/odd disparity observed in the magnetic properties of CAC($n$)+h-BN are also not seen for BN chains grown on graphene. Single B ad-atom as well as single N ad-atom adsorbed on graphene have magnetic ground states. However, while BN(2) and NB(2) chains on graphene are nonmagnetic, BN(3) and NB(3) on graphene have magnetic ground states. This order is disturbed by the chain consisting of four atoms; while BN(4)+graphene is nonmagnetic, NB(4)+graphene has magnetic ground state. The electronic structures calculated for a single adsorbate, namely single boron and nitrogen ad-atoms, and single NB(2) and BN(2) molecules, perpendicularly attached to (4x4) supercell of graphene are shown in Fig.~\ref{fig11}. Owing to their odd number of electrons, single N and B ad-atoms shift the Fermi level up and down relative to the Dirac point of graphene. This is an important feature to be used in electronic applications of single layer graphene. For NB(2)+graphene the bands crossing at the Fermi level attribute semimetallic character. On the other hand, BN(2)+graphene is a semiconductor with a narrow indirect band gap. This situation shows the dramatic effect of the type of chain atom which is attached to graphene. In addition, the supercell size, adsorbates-adsorbate coupling, as well as the symmetry of the adsorbate+supercell system are crucial for the resulting electronic structure for small supercells.\cite{hasanmesh,lambin} For example, the semiconductor presented in Fig.~\ref{fig11} (d) becomes metallic when BN+graphene system is treated in (6x6) supercell. In the case of very large supercell having single adsorbate attached the band crossing is recovered and the states of adsorbate appear as localized (resonance) impurity states. \section{Conclusion} In conclusion, using first-principles calculations within the density functional theory, we revealed the growth mechanism of carbon and BN atomic chains on single layer honeycomb structures, namely hexagonal BN(or h-BN) and graphene, respectively. We found that with the inclusion of each new carbon atom, the existing atomic chain consisting of $n$ carbon atom leaves its previous position to join with the new carbon atom to from a chain of $n+1$ carbon atom. Similar, but more complex growth mechanisms were also found for BN chains grown on graphene layers depending on the type of chain atom(B or N) which is attached to graphene. These growth processes were simulated by ab-initio molecular dynamics calculations at various temperatures and the resulting chain structures were shown to be stable even though they are free to bend and slightly tilt. Nevertheless, these simulations are performed at high temperatures(1000K) and the chains did not detach from the substrate plane for 10ps which is adequately long for ab-initio molecular dynamics simulations. Therefore these simulations presents reliable stability results, especially at room temperature. The growth of atomic chains on the single layer honeycomb structures heralds a self-assembly process, which may have fundamental and technological implications. The grown chains by themselves, exhibit interesting physical and chemical properties depending on the number of atoms forming the chain and the type of chain atom attaching to the substrate. In particular, the physical properties of even numbered and odd numbered carbon chains can behave differently leading to interesting even-odd disparity. We also showed that atomic chains grown on single layer substrates attribute useful functionalities to bare h-BN and graphene. These properties are dependent on the number of carbon atoms in the chains. Apart from creating localized electronic states in the band gap and local magnetic moments it is demonstrated that carbon chains can also be used for increasing the interlayer spacing between BN flakes, where specific molecules can be stored. Atomic chains grown on graphene and h-BN create chemically active sites on BN surface for atoms such as Au, Li and H$_2$, which provides the connection of these materials. Additionally, specific atoms forming strong bonds with atomic chains modify the electronic structure in the band gap and hence change the conductivity for possible sensor applications. Similarly, modifications of the electronic properties of graphene depending on the grown B or N ad-atoms and BN or NB chains are also worth emphasizing. Moving the Fermi level of graphene up or down by adsorbing N or B atoms, respectively can be a worthwhile feature in the electronic applications related with graphene. The self-assembly in terms of atomic chains grown on semimetallic graphene and semiconducting h-BN and interesting functionalities achieved thereof can bring up new perspectives for further research on single layer honeycomb structures. \section{Acknowledgements} Part of the computational resources has been provided by TUBITAK ULAKBIM, High Performance and Grid Computing Center (TR-Grid e-Infrastructure) and UYBHM at Istanbul Technical University through Grant No. 2-024-2007. This work was supported partially by the Academy of Sciences of Turkey(TUBA).
\section{Introduction} For over two decades it has been possible to use numerical methods to model systems of cold, collisionless dark matter particles collapsing under gravity to form stable, virialized `halos' (\citeauthor{1985Natur.317..595F} 1985; \citeauthor{1991ApJ...378..496D} 1991; for a review see \citeauthor{2012arXiv1210.0544F} 2012). The density of these halos declines with radius following a slowly changing power law dependence, roughly $\rho \sim r^{-1}$ at small radii and $\rho \sim r^{-3}$ in the outer regions \citep{1996ApJ...462..563N,1997ApJS..111...73K,1998ApJ...499L...5M}. Despite some early uncertainty, recent simulations with independent computer codes all reproduce this result \cite[e.g.][]{2008Natur.454..735D,2010MNRAS.402...21N,2009MNRAS.398L..21S}. The universal behaviour seems to be independent of the power spectrum of initial linear density fluctuations \citep{1999MNRAS.310.1147M,2005MNRAS.357...82R,2009MNRAS.396..709W} as well as the mass of the collapsed object and the epoch of collapse, although together these determine a scale radius for the transition from $r^{-1}$ to $r^{-3}$ behaviour \citep{1996MNRAS.281..716C,1997ApJ...490..493N,2001MNRAS.321..559B,2001ApJ...554..114E,2007MNRAS.378...55M}. Even simulations of `cold collapse', for which initial conditions consist of a homogeneous sphere of particles, seem to produce similar universal profiles \citep{1999ApJ...517...64H}. It remains an outstanding question, however, whether this universality can be adequately explained from first principles. Until this question is answered, we do not fully understand what the universality means and must rely on new simulations to predict the effect of changes in the initial conditions or particle properties. The experimental result that monolithic collapse produces the same types of system as hierarchical merging \citep{1999ApJ...517...64H,1999MNRAS.310.1147M,2009MNRAS.396..709W} is provocative: it means that any explanation for universality which invokes a specific cosmology \citep[e.g.][]{1998MNRAS.293..337S,2003ApJ...588..680D,2012MNRAS.423.2190S} must be describing a special case of a more general process \citep{2003ApJ...593...26M}. Attempts to understand collisionless gravitational collapse's insensitivity to initial conditions were pioneered by \cite{1967MNRAS.136..101L} in the context of self-gravitating stellar systems. Adopting Boltzmann's procedure for deriving the thermodynamics of collisional systems, Lynden-Bell maximized the entropy of the systems subject to fixed energy. This implies a density profile obeying $\rho(r) \propto r^{-2}$, so disagrees with the results of numerical experiments. Moreover this approach gives rise to a number of physically questionable conclusions \citep[for reviews see][]{1990PhR...188..285P,1999PhyA..263..293L}. The clearest of these is the `gravothermal catastrophe': entropy can be increased without bound by transferring energy from the innermost orbits of a self-gravitating system to the outermost orbits \citep{1968MNRAS.138..495L,1986MNRAS.219..285T}. This implies the existence of a runaway physical instability in which the majority of material collapses into an extreme central density cusp or black hole. Observations and numerics both suggest that something prevents the above catastrophe from occurring on any reasonable timescale; in other words, a physical constraint is preventing the arbitrary redistribution of energy \citep{1987MNRAS.229..103W}. In the spirit of \cite{1957PhRv..106..620J} a general explanation for the final state can still be based on the ideas of statistical mechanics but in the presence of constraints other than energy. The additional constraints will represent the incompleteness of the energy equilibriation effects of violent relaxation. This is the approach we adopt in the present work. The likely distribution of particles in phase space is selected by maximizing entropy, \begin{equation} S= -k \int f\, \ln f\, \mathrm{d}^6 \omega\textrm{,}\label{eq:entropy} \end{equation} where $k$ is Boltzmann's constant and $f(\omega)$ is the probability of finding a particle in a specified region of phase space $\omega$, subject to relevant constraints on the desired solution. (This approach can be motivated by showing that the vast majority of states consistent with a given set of constraints are to be found near the maximum entropy solution; see Appendix \ref{sec:but-what-about} for further discussion and references.) The constraints applied arise from the dynamical evolution of collisionless systems. We will argue that, in the late stages of violent relaxation, there is a diffusion of particles in phase space which approximately conserves the sum of orbital actions. This sum is progressively better conserved as equilibrium is approached and therefore its role in establishing that equilibrium cannot be ignored. Applying the maximum entropy recipe, we will show that the phase space structure of equilibrium halos is then reproduced over orders of magnitude in probability density. To our knowledge, this is the first instance of maximum entropy reasoning, applied to 6D phase space and subject to physically motivated constraints, producing such success in a collisionless system. The remainder of this work is organized as follows. First, the conservation of action is discussed (Section \ref{sec:conservation-action}). Then a canonical ensemble constructed on this basis (Section \ref{sec:new-canon-ensemble}) yields a phase space distribution in quantitative agreement with high resolution numerical experiments (Section \ref{sec:simulations}). A discrepancy affecting a small fraction of particles at low angular momentum is highlighted in Section \ref{sec:results-j}. Finally we discuss the predicted radial density profiles which again highlight the need for special treatment of low angular momentum orbits (Section \ref{sec:real-space}). We conclude in Section \ref{sec:conclusions}. \section{The analytic phase space distribution} \subsection{Conservation of action}\label{sec:conservation-action} Our first task is to identify and explain some relevant quantities which should be held fixed when maximizing entropy. This will be central to our argument because such constraints represent the incompleteness of violent relaxation's tendency to redistribute energy, generating a different solution from the one based on energy conservation alone. With this in mind we will show that the radial action $J_r$ (to be defined below) is conserved in an average sense even during rapid potential changes. This average conservation does not appear to have been discussed elsewhere in the literature. We will first show how it can be derived from previous work \citep{2012MNRAS.421.3464P} when the potential changes instantaneously, maintaining the sphericity of the halo. Then a more general (but more abstract) argument will be given which additionally shows that the other two actions (the $z$-component of the angular momentum $j_z$ and the scalar angular momentum $j$) are also conserved in the same average sense. The second approach encompasses perturbations to the potential which have variations on arbitrary timescales and may break spherical symmetry. However the first has a more intuitive content and therefore forms our starting point. In a spherical system, the radial action $J_r$ is defined by \begin{equation} J_r = \frac{1}{\pi} \int_{r_{\mathrm{min}}}^{r_{\mathrm{max}}} \sqrt{2E-2\Phi(r; t) - j^2/r^2}\, \mathrm{d} r\textrm{,}\label{eq:radial-action} \end{equation} Here $E$ is the specific energy, $j$ is the specific angular momentum and $\Phi$ is the potential at a given radius $r$ and time $t$; the $r$ integral is taken over the region where the integrand is real. The radial action $J_r$ has the same units as specific angular momentum $j$. This reflects the similar conservation roles these two quantities play for the radial and angular components of the motion. In particular $J_r$ is exactly conserved if any changes in the potential occur sufficiently slowly (`adiabatically') in time \citep[e.g.][]{1987gady.book.....B}. On the other hand in the rapid, impulsive limit under an instantaneous change in energy $E \to E + \Delta E$ and potential $\Phi(r) \to \Phi(r) + \Delta \Phi(r)$, the action of the particle is changed at first order: \begin{equation} \Delta J_r = \left.\frac{\partial J_r}{\partial E}\right|_{\Phi} \Delta E + \int_0^{\infty} \mathrm{d} r\, \Delta \Phi \left. \frac{\delta J_r}{\delta \Phi} \right|_{E}\textrm{.} \label{eq:delta-J-delta-E-relation} \end{equation} In \cite{2012MNRAS.421.3464P} we showed that the energy shift $\Delta E$ induced by the change of potential, averaged over possible orbital phases of the particle, is \begin{equation} \langle \Delta E \rangle = - \frac{\int_0^{\infty} \mathrm{d} r \,\Delta \Phi \left. \delta J_r / \delta \Phi \right|_{E}}{\left. \partial J_r / \partial E \right|_{\Phi}}\textrm{,}\label{eq:average-energy-shift} \end{equation} an exact result \cite[see equation 12 of][]{2012MNRAS.421.3464P}. Here angular brackets denote averaging over all possible phases of the orbit. Considering the probability distribution of radial actions after this change, one has $\langle \Delta J_r \rangle = 0$ at first order, by substituting equation \eqref{eq:average-energy-shift} in \eqref{eq:delta-J-delta-E-relation}. Even though a specific particle will change its radial action, the ensemble average is conserved. This result connects closely with the standard adiabatic argument that $\Delta J_r=0$ if any changes to the potential occur on long timescales. In our case, however, the necessary `phase averaging' does not occur over time for an individual particle but instead via a statistical consideration of an ensemble of particles spread evenly through all possible phases. \begin{figure} \includegraphics[width=0.49\textwidth]{diffuse.pdf} \caption{An illustration of the diffusion of particles in action space. Here particles in the inner $10\,\mathrm{kpc}$ of a simulation have been selected and their radial actions $J_r$ numerically calculated at two timesteps separated by $\Delta t=2.7\,\mathrm{Gyr}$. The change in the population mean action is small ($\mu = -12.1\kpc\,\mathrm{km\,s^{-1}}$) compared against the magnitude of the random diffusion ($\sigma =287\,\kpc\,\mathrm{km\,s^{-1}}$). }\label{fig:diffuse} \end{figure} We can generalize as follows. Adopting the complete set of action-angle coordinates for phase space \citep[e.g.][]{1979LNP...110.....P,1987gady.book.....B}, the momenta are $\vec{J}=(J_r, j, j_z)$ where $j$ is the total angular momentum and $j_z$ is its component in the $z$ direction (so $-j<j_z<j$). The conjugate coordinates are $\vec{\Theta}=(\psi_r, \phi, \chi)$, taken to be periodic with interval $2 \pi$, and the Hamiltonian is \begin{equation} H(\vec{J}, \vec{\Theta}) = H_0(\vec{J}, t) + h(\vec{J}, \vec{\Theta}, t)\textrm{.} \end{equation} Here $h$ is an arbitrary perturbation. It may consist of a long-lived term (perhaps a departure from spherical symmetry) and fluctuations on arbitrary time-scales. In the background, the $\vec{\Theta}$ coordinates change at a constant rate, \begin{equation} \vec{\Theta} = \vec{\Theta}_0 + \vec{\Omega} \, t\textrm{,}\label{eq:theta-of-t} \end{equation} and the frequencies for the radial and azimuthal motion $\Omega_r$ and $\Omega_j$ obey \begin{equation} \Omega_r = \frac{\partial{H_0}}{\partial J_r}; \hspace{0.5cm} \Omega_j = \frac{\partial H_0}{\partial j},\label{eq:omegas} \end{equation} by Hamilton's equations. The frequencies may change slowly with time ($\mathrm{d} \Omega /\mathrm{d} t \ll \Omega^2$). $\vec{J}$ is conserved in the background but when $h$ is non-zero, the equations of motion read \begin{equation} \frac{\mathrm{d} \vec{J}}{\mathrm{d} t} = - \frac{\partial{h}}{\partial\vec{\Theta}}\textrm{.} \end{equation} Because equation \eqref{eq:theta-of-t} shows that particles in the background move at a uniform rate in the $\vec{\Theta}$ coordinates, an equilibrium distribution $f$ has no $\vec{\Theta}$ dependence; {\it i.e.} the density $f$ is a function of $\vec{J}$ alone\footnote{The conjugate coordinate to $j_z$ is also a constant of motion in the background, so this argument does not strictly show $\mathrm{d} \langle j_z \rangle/\mathrm{d} t =0$. However $\langle j_z \rangle$ must be exactly conserved anyway if the perturbations are internally generated, since it is proportional to a component of the total angular momentum vector.}. From this it follows that \begin{equation} \frac{\mathrm{d}}{\mathrm{d} t} \langle \vec{J} \rangle = -\int \mathrm{d}^3 J \mathrm{d}^3 \Theta f(\vec{J}) \frac{\partial h}{\partial \vec{\Theta}} = 0\textrm{,} \end{equation} where the result is obtained via integration by parts. This means an individual particle's $\vec{J}$ can `diffuse' \citep{1988MNRAS.230..597B} over large distances $\sigma$ in action space, \begin{equation} \sigma \equiv \langle (\delta J)^2 \rangle^{1/2} = \mathcal{O}(\epsilon)\textrm{,} \end{equation} compared to the variation in the mean $\mu$ of the population, \begin{equation} \mu \equiv \langle \delta J \rangle = \mathcal{O}(\epsilon^2). \end{equation} We can inspect this diffusion in a simulation by calculating the relevant actions at two timesteps. As an example, Figure \ref{fig:diffuse} shows a histogram of changes in the radial action of tightly bound particles in the forming ``Dwarf'' halo (see Section \ref{sec:simulations}) between $z=3.1$ and $z=1.4$, a time interval of approximately $2.7\,\mathrm{Gyr}$. We define `tightly bound' by selecting particles interior to $10\,\mathrm{kpc}$ at the earlier time step, and calculate the $J_r$ values of these particles in both outputs according to the numerical recipes given later (Section \ref{sec:perf-comp-techn}). As expected from the linear analysis, Figure \ref{fig:diffuse} shows that individual simulated particles change their actions more rapidly than the population mean. Quantitatively, the change in the population mean $\mu$ is $-12.1\,\kpc\,\mathrm{km\,s^{-1}}$ whereas a typical particle has moved by $\sigma=287\,\kpc\,\mathrm{km\,s^{-1}}\gg |\mu|$ from its original $J_r$ value. Note also that the mean $J_r$ value for these particles in the final timestep is $\langle J_r \rangle = 187\,\kpc\,\mathrm{km\,s^{-1}}<\sigma$, meaning particles really do cover significant distances in action space. This analysis confirms that $\langle \vec{J} \rangle$ evolves slowly and, although it is not exactly conserved, it forms a constraint of motion that cannot be ignored on finite timescales. A complete description would require investigation of different moments of the distribution at higher order in perturbation theory. For now, however, we have motivated a picture in which $\langle \vec{J} \rangle$ evolves sufficiently slowly compared to the diffusion of an individual particle that it must be considered fixed in analysis of the distribution. \subsection{The new canonical ensemble}\label{sec:new-canon-ensemble} In the Introduction we explained that, to obtain a phase space distribution function, we will maximize the entropy~\eqref{eq:entropy} subject to constraints on the particle population (further discussion is given in Appendix \ref{sec:but-what-about}). As well as energy conservation, we apply the 3-vector of constraints on $\langle \vec{J} \rangle$ discussed above. This gives rise to a total of four Lagrange multipliers in the resulting distribution function: \begin{equation} f(\vec{J}) \propto \exp \left(-\vec{\beta} \cdot \vec{J} -\beta_E E(\vec{J})\right)\textrm{,}\label{eq:maxent-solution} \end{equation} where the Lagrange multipliers are $\vec{\beta} = (\beta_r, \beta_j, \beta_z)$ and $\beta_E$. In the absence of the new constraints, $\vec{\beta}=0$ and $\beta_E$ is identified with $1/kT$ (where $T$ is the thermodynamical temperature). All four constants can be determined in a variety of ways depending on the situation; for a complete account of structure formation one would like to be able to derive them from the initial conditions, but this lies beyond the scope of the current paper (although see Section \ref{sec:other-simulations} for further comments). The lack of any reference to $\vec{\Theta}$ in equation~\eqref{eq:maxent-solution} indicates that the solution is phase-mixed, as required for equilibrium. Equation~\eqref{eq:maxent-solution} is the essential prediction of the present work. As with any prediction derived from a maximum entropy argument, it will be able to fit the actual ensemble only if we have encapsulated enough of the dynamics within the constraints \citep{Jaynes79}. The rest of this paper is concerned with testing to what extent that is the case. \begin{figure*} \begin{center} \includegraphics[width=18cm]{dm_images.pdf} \begin{tabularx}{17.8cm}{XXX} \textbf{Dwarf} & \textbf{MW} & \textbf{Cluster} \\ $r_{200} = 98\,\mathrm{kpc}$; $M_{200}=2.8 \times 10^{10}\,\mathrm{M}_{\odot}$; $c=19.6$ & $r_{200} = 301\,\mathrm{kpc}$; $M_{200}=8.0 \times 10^{11}\,\mathrm{M}_{\odot}$; $c=15.5$ & $r_{200} = 1.43\,\mathrm{Mpc}$; $M_{200} = 8.7\times10^{13}\,\mathrm{M}_{\odot}$; $c=9.9$ \\ $N_{\mathrm{part}} = 3.4 \times 10^6$; $\epsilon = 65\,\mathrm{pc}$. & $N_{\mathrm{part}} = 5.3 \times 10^6$; $\epsilon = 170\,\mathrm{pc}$. & $N_{\mathrm{part}} = 8.9 \times 10^6$; $\epsilon = 690\,\mathrm{pc}$. \end{tabularx} \end{center} \caption{Images of the three dark matter simulations, accompanied by numerical properties. Respectively $r_{200}$, $M_{200}$, $c$, $N_{\mathrm{part}}$ and $\epsilon$ denote the radius at which the density exceeds the critical density by a factor $200$; the mass within this radius; the `concentration', $c=r_{200}/r_s$ where $r_s$ is the NFW scale radius as described in the text; the number of particles within $r_{200}$; and the gravitational softening length in physical units at $z=0$. The images are scaled to show the virial sphere of the main halo. The brightness represents the column density of dark matter (scaled logarithmically to give a dynamic range of $3000$ in each case); the colour corresponds to a density-weighted potential along the line of sight. }\label{fig:simulations} \end{figure*} First consider the probability of finding a particle with $J_r$ in a given interval (ignoring $j$ and $j_z$ coordinates). This is given by \begin{equation} p_r(J_r) = \int_0^{2\pi} \mathrm{d}^3 \Theta \int_0^{\infty} \mathrm{d} j \int_{-j}^j \mathrm{d} j_z\, f(\vec{J})\textrm{,}\label{eq:marginalize-out-j} \end{equation} because the action-angle coordinates are canonical, so the phase-space measure is constant. In the limit that the energy of the system becomes large at fixed action, equation \eqref{eq:marginalize-out-j} can be solved: \begin{equation} p_r(J_r) \propto \exp - \beta_r J_r\hspace{1cm}(\beta_E=0)\textrm{,}\label{eq:Jr-only-distrib} \end{equation} but it is not immediately clear whether we will be operating in this regime. More generally a closed form for $p_r(J_r)$ is hard to obtain, but we can at least show that \begin{equation} \frac{\mathrm{d} \ln p_r}{\mathrm{d} J_r} = -\beta_E \langle \Omega_r \rangle_{J_r} - \beta_r\textrm{,}\label{eq:MAXENT-Jr-derivative} \end{equation} which can then be integrated numerically for a given case to give a concrete comparison between equation \eqref{eq:maxent-solution} and simulations. Here we have defined \begin{equation} \langle \Omega_r \rangle_{J_r} \equiv \frac{1}{p_r(J_r)}\int f(\vec{J}) \Omega_r(\vec{J})\, \mathrm{d} j\, \mathrm{d} j_z\textrm{,} \end{equation} which is the mean of the radial frequency for particles with a fixed $J_r$. With this definition, relation \eqref{eq:MAXENT-Jr-derivative} can be derived from equation \eqref{eq:marginalize-out-j}, recalling that the radial frequency $\Omega_r$ of the particle's orbit obeys equation~\eqref{eq:omegas}. We will investigate and explain the distribution of $J_r$ values predicted by equation~\eqref{eq:MAXENT-Jr-derivative} in Section \ref{sec:results}. Now consider the distribution of total angular momentum $j$. We will follow exactly the same series of manipulations as for the radial action; however in this case the marginalization over $j_z$ introduces a non-trivial term: \begin{align} p_j(j) & = \int_0^{2 \pi} \mathrm{d}^3 \Theta \int_{-j}^{j} \mathrm{d} j_z\, \int_0^{\infty} \mathrm{d} J_r\, f(\vec{J}) & \nonumber \\ & \propto \sinh \left( \beta_z j \right) \, \exp \left(-\beta_j j \right) \int_0^{\infty} \mathrm{d} J_r \exp \left( - \beta_E E \right) \textrm{.}\label{eq:pj-function} \end{align} Once again, in the case $\beta_E\to 0$, we have a fully analytic expression for $p_j(j)$\textrm{,} \begin{equation} p_j(j) \propto \sinh \left(\beta_z j \right) \exp \left( - \beta_j j \right)\hspace{1cm}(\beta_E=0)\textrm{,}\label{eq:angmom-no-E} \end{equation} which will serve as a useful point of comparison. More generally we can differentiate equation \eqref{eq:pj-function} to obtain \begin{equation} \frac{\mathrm{d} \ln p_j}{\mathrm{d} j} = \beta_z \coth \left(\beta_z j\right) -\beta_E \langle \Omega_j \rangle_j - \beta_j\textrm{,}\label{eq:MAXENT-angmom} \end{equation} where $\Omega_j$ is the angular frequency of the orbit and \begin{equation} \langle \Omega_j \rangle_j \equiv \frac{1}{p_j(j)} \int f(\vec{J}) \Omega_j(\vec{J}) \mathrm{d} J_r \, \mathrm{d} j_z \end{equation} is the mean angular frequency of particles at fixed $j$. Equation \eqref{eq:MAXENT-angmom} for the angular momentum distribution (ignoring all other coordinates) is the equivalent of equation \eqref{eq:MAXENT-Jr-derivative} for the radial action distribution. Once again we will investigate and explain the shape it predicts in Section~\ref{sec:results-j}. First, however, we will explain the simulations which serve as a point of comparison for the later discussions. \section{Comparison to simulations}\label{sec:simulations} \subsection{Overview of the simulations} \begin{figure*} \includegraphics[width=0.9\textwidth]{h285_Jr.pdf} \caption{The distribution of particles' radial action $J_r$ (left panel) and scalar angular momentum $j$ (right panel) in MW (both panels show the distribution of particles as a histogram). Also shown are maximum entropy solutions based on energy conservation alone (dotted curve); action conservation alone (dashed curve) and our advocated solution using both constraints (thick solid curve). The last of these provides a good reproduction of the distribution for $J_r<J_{r,\mathrm{break}} = 3 \times 10^4\,\mathrm{kpc\,km\,s^{-1}}$ and $j<j_{\mathrm{break}}=4\times 10^4\,\mathrm{kpc\,km\,s^{-1}}$ respectively, while extending over orders of magnitude in probability density. Less than $0.1\%$ of particles lie at $J_r>J_{r,\mathrm{break}}$; approximately $0.2\%$ lie at $j>j_{\mathrm{break}}$. Despite the good overall agreement, problems become apparent at very small $j$; a blowup of the indicated range $j<3500\, \kpc\,\mathrm{km\,s^{-1}}$ is given in Figure 4. }\label{fig:Jr} \end{figure*} In the previous section we applied maximum entropy reasoning to conservation of energy and approximate conservation of action to derive an expected equilibrium phase space distribution. We will now compare that expectation against simulated dark matter halos. Our strategy is to integrate equations \eqref{eq:MAXENT-Jr-derivative} and \eqref{eq:MAXENT-angmom} numerically for these simulations and compare to the actual distribution of particles binned by $J_r$ and $j$ respectively. We will present results from three simulated dark matter halos (shown in Figure~\ref{fig:simulations}), chosen to span a wide range of masses with an approximately constant number of particles per halo (several million in each case). We also compared our results against the GHALO multi-billion-particle phase space \citep{2009MNRAS.398L..21S}, finding good agreement similar to that described for our ``MW'' halo here. This gives confidence that the mechanisms and results discussed in the paper are not sensitive to numerical resolution. Our simulations are run from cosmological initial conditions at $z\simeq 100$ in a `zoom' configuration \citep{1993MNRAS.265..271N}, i.e. with high resolution for the main halo and its immediate surroundings and lower resolution for the cosmological environment. The softening lengths $\epsilon$ for the high resolution region are listed in Figure~\ref{fig:simulations} and are fixed in physical units from $z=9$, prior to which they scale linearly with cosmological scalefactor, a compromise motivated by numerical convergence studies \citep{2004MNRAS.348..977D}. We verified at the final output ($z=0$) that the high resolution regions have not been contaminated by low resolution particles, and that the halo real-space density profiles are well described by a slowly rolling powerlaw, in accordance with all recent simulations \citep[e.g.][and references therein]{2008Natur.454..735D,2009MNRAS.398L..21S,2010MNRAS.402...21N}. Each simulation output contains full cartesian phase space coordinates ($\vec{x}$, $\vec{v}$). The position space is re-centred on the central density peak of the halo using the `shrinking sphere' method of \cite{2003MNRAS.338...14P}. The velocities are re-centred such that a central sphere of radius $r_{200}/30$ has zero net velocity, where $r_{200}$ is the radius at which the mean halo density is $200$ times the critical density. Henceforth we only consider particles inside $r_{200}$. From left to right in Figure \ref{fig:simulations} the simulated halos become more massive. The width of each panel is equal to $2r_{200}$ and the luminosity is scaled to represent the column density over a dynamic range of $3000$. The most conspicuous aspect of Figure 2 is that the halos become less centrally concentrated. We verified this by fitting a classic ``NFW'' \citep{1996ApJ...462..563N} formula to the density profile. The NFW fit, \begin{equation} \rho(r) = \frac{\rho_0}{(1+r/r_s)^2\,(r/r_s)}\textrm{,} \end{equation} yields $\rho_0$, a characteristic density, and $r_s$, a scale radius. The latter is often expressed in a scale-free manner as a concentration value $c=r_{200}/r_s$; we have recorded the value for each halo in Figure \ref{fig:simulations}. As expected the concentration decreases with increasing mass, in agreement with previously known trends \citep[e.g.][]{2007MNRAS.378...55M,2001MNRAS.321..559B}. We thus have a sample of cosmological halos which span a wide range in both mass and concentration. These different concentrations are thought to arise from different mean densities in the universe at the epoch of collapse \citep{1997ApJ...490..493N,2001MNRAS.321..559B}. \label{sec:perf-comp-techn} All halos in Figure \ref{fig:simulations} exhibits large amounts of substructure; we will present results with this substructure subtracted, although we have verified that including the substructure does not have a qualitative impact on our results. The substructure is identified and removed using the ``Amiga Halo Finder'' \citep{2009ApJS..182..608K}. For each remaining particle inside $r_{200}$, the specific scalar angular momentum is given by $j=\left|\vec{v} \times \vec{x}\right|$. We calculate $J_r$ by evaluating equation~\eqref{eq:radial-action} numerically, using a spherically-averaged potential $\Phi$ defined by \begin{equation} \Phi(r) = \int_0^r \mathrm{d} r' \frac{GM(<r')}{r'^2}\textrm{,}\label{eq:Phi-equation} \end{equation} where $M(<r')$ is the total mass enclosed by a sphere of radius $r'$, and the specific energy $E$ of each particle is defined as $E=\vec{v}^2/2 + \Phi(r)$. $J_r$ is evaluated using the true spherical potential out to $r_{\mathrm{term}} = 3 r_{200}$, beyond which (for reasons of numerical speed) the calculation is truncated and an analytic completion assuming a Keplerian (vacuum) potential is taken. We verified that changing $r_{\mathrm{term}}$ to $4 r_{200}$ had little impact on the results. Before proceeding to a comparison, we need to derive appropriate $\beta$ values. We calculate these using a Monte-Carlo Markov chain (MCMC) to maximize the likelihood \begin{equation} \mathcal{L}(\vec{\beta}, \beta_E) = \prod_i f(\vec{J}_i; \vec{\beta}, \beta_E) \end{equation} where $f$ is the 1-particle distribution function \eqref{eq:maxent-solution} normalized such that $\int \mathrm{d}^3 \vec{J}\, \mathrm{d}^3 \vec{\Theta} \,f(\vec{J}) = 1$. This normalization must be accomplished numerically on a grid of $J_r, j$ values; at each grid-point $E(J_r, j)$ is calculated by operating a bisection search on equation \eqref{eq:radial-action}. This need only be done once, and then the evaluation of each link in the Markov chain is rapid. The operation gives us maximum likelihood (i.e. ``best fit'') parameters\footnote{The MCMC technique also yields uncertainties on the $\beta$ values, but these will not be considered further in the present work.} $(\beta_j, \beta_z, \beta_r, \beta_E)$ for a given simulation, optionally subject to constraints (such as $\beta_E=0$ or $\beta_j = \beta_z=\beta_r=0$). We are now fitting up to four parameters (excluding mass normalization), more than the one or two parameters normally used by simulators to describe their halos \citep[e.g.][]{2004MNRAS.349.1039N,2009MNRAS.398L..21S}. However the fitted real-space density profiles are purely phenomenological constructs; conversely here we are starting with a functional form derivable from physical considerations. As we have commented in Section \ref{sec:new-canon-ensemble} and will expand upon in Section \ref{sec:other-simulations}, the $\beta$'s should ultimately therefore be derived from initial conditions. For the present, however, the objective is to see whether our physical argument can correctly describe the phase space distribution at all, for which fitting $\beta$'s is the most pragmatic approach. \subsection{Comparison with MW: $J_r$ distribution}\label{sec:results} We will now start to test how closely equation \eqref{eq:maxent-solution} represents the distribution of particles in our simulations. We will investigate MW in some detail, before showing results for the other two simulations to which the same discussion can essentially be applied. The distribution of $J_r$ values in MW is shown by the histogram in the left panel of Figure \ref{fig:Jr}. This can be compared with the thick solid curve which shows the distribution of $J_r$ values according to expression \eqref{eq:MAXENT-Jr-derivative}; the parameters are $\beta_E^{-1} = 1.6 \times 10^4 \,\mathrm{km^2\,s^{-2}}$ and $\beta_r^{-1} = 4.4 \times 10^3 \,\kpc\,\mathrm{km\,s^{-1}}$. The agreement is excellent over several orders of magnitude in probability density, spanning the values $0<J_r<J_{\mathrm{break}}$ where $J_{\mathrm{break}} \simeq 3 \times 10^4\, \kpc\,\mathrm{km\,s^{-1}}$. Particles with $J_r>J_{\mathrm{break}}$ account for less than $0.1\%$ of the mass and are on long period orbits, probably reflecting new material falling into the potential well. We will not consider them further. The shape of the $J_r$ solution can be understood as follows. We have already remarked that, in the limit $\beta_E \to 0$, one recovers equation \eqref{eq:Jr-only-distrib}, an exact exponential (i.e. a straight line on the linear-log axes of Figure \ref{fig:Jr}). For comparison we have plotted the best fit distribution of this form (with $\beta_r^{-1} = 3.5 \times 10^3\, \kpc\,\mathrm{km\,s^{-1}}$) as a dashed line. Since the period of an orbit increases with its energy (or radial action), the mean frequency $\langle \Omega_r \rangle_{J_r}$ decreases for increasing $J_r$. So, inspecting equation \eqref{eq:MAXENT-Jr-derivative}, there will always be a $J_r$ value above which $\beta_E \langle \Omega_r \rangle_{J_r}$ becomes much smaller than $\beta_r$. Looking again at the thick solid curve in Figure \ref{fig:Jr}, the limiting solution at high $J_r$ is indeed a pure exponential as this reasoning would suggest. At small $J_r$, however, the gradient of the solution is steeper because of the energy term. Comparing the histogram, the thick solid curve and the dashed line in the left panel of Figure \ref{fig:Jr} thus leads us to the conclusion that $J_r$ conservation (dashed line) accounts rather well for the qualitative form of the distribution, with an important correction from $E$ conservation at low $J_r$. Finally the dotted curve shows the best fit case with $\beta_r=0$ -- i.e. the normal statistical mechanical result in the absence of other constraints -- and provides a poor fit at all $J_r$. In summary, the identification of the $J_r$ constraint has resulted in dramatic improvements in the match to simulations. \subsection{Comparison with MW: $j$ distribution}\label{sec:results-j} \begin{figure} \includegraphics[width=0.47\textwidth]{h285_j_blowup.pdf} \caption{Despite good agreement over the majority of $j$ space (see right panel of Figure \ref{fig:Jr}), the fraction of simulated orbits (histogram) at very low angular momentum is substantially underestimated by the simplest maximum entropy argument (thin dotted curve). One fix discussed in the text is to postulate a second population at low energies (dashed curve). This yields a much better low-$j$ fit (solid line) without affecting the high-$j$ fit (except through a minor renormalization). }\label{fig:h285-blowup} \end{figure} \begin{figure*} \includegraphics[width=0.9\textwidth]{other_distribs.pdf} \caption{As Figure \ref{fig:Jr}, but for the remaining two simulations. Once again, the maximum entropy distribution subject to $\vec{J}$ and $E$ constraints (thick solid lines) predicts the simulations (histogram) accurately. The simulated distribution in $J_r$ for the cluster (lower left panel) has some noticable fluctuations over large scales; this is likely because it is dynamically young.}\label{fig:others} \end{figure*} Now consider the right panel of Figure \ref{fig:Jr} which shows the distribution of scalar angular momentum for the particles in our MW simulation. Once again the simulated particles are shown by the histogram; the best fit maximum entropy solution ($\beta_j^{-1}, \beta_{z}^{-1} = 4.4\times 10^3, 1.1\times 10^4\, \kpc\,\mathrm{km\,s^{-1}}$, with $\beta_E$ as quoted above) is shown by the thick solid curve. It again reproduces the correct qualitative behaviour up to $j_{\mathrm{break}} = 4 \times 10^4 \,\kpc\,\mathrm{km\,s^{-1}}$, with only $0.2\%$ of the mass at $j>j_{\mathrm{break}}$. Although the angular momentum distribution has some fluctuations away from the predicted behaviour, the predictions remain nearly correct over two orders of magnitude in probability density. With the exception of a problem described below, we do not believe these fluctuations to be of particular importance beyond indicating the structure is not completely relaxed. In particular we will show later (Section \ref{sec:real-space}) that these inhomogeneities can be ignored when reconstructing a density profile in real space. Certainly compared against a solution based on $E$ conservation alone, again shown by a dotted curve, our solution can be counted a success. The basic shape of the predicted $j$ distribution can be understood in a similar way to the $J_r$ distribution explained above. Consider again the case where $\beta_E=0$ (so in effect the total energy is unconstrained); then the exact solution is given by equation~\eqref{eq:angmom-no-E}. We can also take the isotropic limit, $\beta_z \to 0$, giving \begin{equation} p_j(j) \propto j\, \exp(-\beta_j j)\hspace{1cm}(\beta_z=0,\textrm{ }\beta_E=0). \end{equation} This is analogous to the radial action case \eqref{eq:Jr-only-distrib}, but with a degeneracy factor $j$ reflecting the increasing density of available states available as the angular momentum vector grows in size. The result is that the abundance of particles grows linearly with $j$ for $j<\beta_j^{-1}$ and decays exponentially for $j>\beta_j^{-1}$. In light of the above discussion, it is notable that the turnover from growth to decay in $p_j(j)$ occurs at $j$ values much smaller than $\beta_j^{-1}$. There are two ways to accomplish this. The first is to create a highly anisotropic setup, $\beta_z^{-1} \ll j_0$, where $j_0$ is the smallest $j$ value of interest. This packs orbits as much as possible into a single plane, generating a large net angular momentum and destroying the approximate spherical symmetry\footnote{We note in passing that, technically, distribution functions with net angular momentum can nonetheless generate spherical potentials \citep{1960MNRAS.120..204L}.}, but effectively removing the degeneracy in $j$ altogether: \begin{equation} p_j(j) \propto \exp \left[(\beta_z-\beta_j) j\right] \hspace{1cm}(\beta_z^{-1}\ll j, \textrm{ }\beta_E=0)\textrm{.} \end{equation} Because it is maximally anisotropic, this solution cannot reflect the simulations; however if we temporarily fit only $j$ values using the functional form \eqref{eq:angmom-no-E}, we are pushed towards this unphysical limit (dashed line, Figure \ref{fig:Jr}, right panel; $(\beta_j^{-1},\,\beta_z^{-1}) = (5.9,\, 6.3)\times 10^2\,\kpc\,\mathrm{km\,s^{-1}}$). Luckily this is not the only way to overcome the shrinking phase space at low $j$. Equation \eqref{eq:MAXENT-angmom} shows that if $\langle \Omega_j\rangle_j$ increases fast enough as $j\to 0$ it can overcome the $\coth(\beta_z j)$ degeneracy term. We have verified that in the full solution (thick solid line in Figure \ref{fig:h285-blowup} right panel), this is the mechanism by which the turnover is pushed to low $j$. Focussing attention on the low-$j$ part of the distribution does, however, reveal a deficiency in our predictions. Figure \ref{fig:h285-blowup} shows the distribution of orbits with $j<3500\,\kpc\,\mathrm{km\,s^{-1}}$. The dotted line shows the same maximum entropy fit depicted by the solid line in Figure \ref{fig:Jr}. When the horizontal scale is expanded in this way, it becomes clear that the global fit undershoots the simulation values significantly at low $j$. This appears to be a systematic feature of all simulations we have inspected (the three detailed here, GHALO, and various other lower resolution simulations which we used for testing purposes). It is possible to force a better fit by restricting the likelihood analysis to this region, but the global agreement is then considerably worse. This suggests that the behaviour at low $j$ is marginally decoupled from that in the rest of phase space. This could arise from the wide range of orbital periods: particles with small actions also have periods much shorter than the rest of the halo. The coupling between particles will necessarily be weak if their timescales are very different (since particles on short orbits react adiabatically to fluctuations on long timescales). This can substantially suppress redistribution of scalar angular momentum and is consistent with, although not reliant on, the early formation of a stable central cusp in simulations \cite[e.g.][]{1998ApJ...499L...5M,2006MNRAS.368.1931L,2011MNRAS.413.1373W}. In principle this weakness of coupling between orbits in different regions could be expressed as a further constraint in the maximum entropy formalism. Further investigation awaits future work, but for now we will use this as a motivation to study a two-population system. We are not suggesting that there really are two sharply defined populations, but that this should anticipate the features of incomplete equilibrium. Our maximum likelihood analysis is able to find a dramatically better fit in this case, placing 3.5\% of the mass in a second population at substantially lower temperature ($\beta_E^{-1} = 2.8\times 10^3 \,\mathrm{km^2\,s^{-2}}$). The summed distribution is shown by the thick solid line in Figure \ref{fig:h285-blowup}, with the contribution from the subdominant population indicated by the dashed line. Because the second distribution is so peaked near $j=0$, the only difference at high $j$ is a marginal renormalization. We also verified that the $J_r$ distribution is barely affected. It is undeniably disappointing that our solution does not automatically accommodate the behaviour at very low $j$, but we expect that future development of the ideas above can quantitatively account for the discrepancy. We consider other possible explanations in Section~\ref{sec:conclusions}. However after focussing so much on one corner of phase space we should re-emphasize the major conclusion: the distribution over both $J_r$ and $j$ for 96\% of the particles are remarkably well described by the maximum entropy expression~\eqref{eq:maxent-solution}. \subsection{Other simulations}\label{sec:other-simulations} We confirmed that these basic conclusions persist in other simulations. The top row of figures in Figure \ref{fig:others} shows results from the `dwarf' simulation. The $J_r$ distribution (top left panel) again shows excellent agreement for $J_r<J_{\mathrm{break}}$, where $J_{\mathrm{break}} = 4\times 10^3 \, \kpc\,\mathrm{km\,s^{-1}}$. Only a tiny fraction of mass ($<0.1\%$) lies beyond this point of breakdown. As with MW, the dwarf's angular momentum distribution (top right panel) has more conspicuous fluctuations, but still roughly adheres to the maximum entropy solution up to $j_{\mathrm{break}}=3 \times 10^3\,\kpc\,\mathrm{km\,s^{-1}}$, with around $1.3\%$ of mass lying beyond this point. We verified that at very low angular momenta $j<100\,\kpc\,\mathrm{km\,s^{-1}}$ there is again an overabundance of particles in the simulation, although in this case it accounts for less than $2\%$ of particles compared against the $3.5\%$ in MW. The parameters of the dwarf fit are $(\beta_r^{-1}, \beta_j^{-1}, \beta_z^{-1}) = (4.2, 1.9, 2.1) \times 10^2\,\kpc\,\mathrm{km\,s^{-1}}$ and $\beta_E^{-1}= 2.2 \times 10^3 \,\mathrm{km\,s^{-1}}$. \begin{figure} \includegraphics[width=0.47\textwidth]{scaling.pdf} \caption{The best fit scales for radial action (upper panel) and energy (lower panel) as a function of mass. The crosses show the values from the three simulations, while dotted lines give the expected scalings \eqref{eq:E-scale} and \eqref{eq:J-scale}, which agree well with the simulations.}\label{fig:scalings} \end{figure} Considering the cluster simulation (lower row of Figure \ref{fig:others}) gives similar results once again. This time the $J_r$ distribution as well as the $j$ distribution shows some notable fluctuations around the maximum entropy description. This may be because clusters assemble later (as we discussed in Section \ref{sec:perf-comp-techn}, this is reflected in the lower concentration value), so the system is dynamically young; however we have not explicitly looked at time dependence of these distributions. The parameters of the cluster fit are $(\beta_r^{-1}, \beta_j^{-1}, \beta_z^{-1}) = (1.5, 0.8, 1.2) \times 10^5\,\kpc\,\mathrm{km\,s^{-1}}$ and $\beta_E^{-1}= 3.5 \times 10^5 \,\mathrm{km\,s^{-1}}$ with $0.1\%$ and $1.1\%$ of the mass in the unrelaxed components beyond $J_{\mathrm{break}} = 7\times 10^5\,\kpc\,\mathrm{km\,s^{-1}}$ and $j=6\times 10^5\,\kpc\,\mathrm{km\,s^{-1}}$ respectively. Naively one would expect $\beta_E^{-1}$ to scale approximately as \begin{equation} \beta_E^{-1} \propto \frac{GM_{200}}{r_{200}} \propto M_{200}^{2/3}\textrm{,}\label{eq:E-scale} \end{equation} since $M_{200}$ and $r_{200}$ are by definition related through the fixed-mean-density condition $M_{200} \propto r_{200}^3$. Similarly the actions $\vec{\beta}^{-1}$ should scale as \begin{equation} \vec{\beta}^{-1} \propto r_{200} \sqrt{\frac{GM_{200}}{r_{200}}} \propto M_{200}^{2/3}\textrm{.}\label{eq:J-scale} \end{equation} Figure \ref{fig:scalings} compares these expectations with the actual values, although we immediately caution against taking the scaling of three halos too seriously. The upper panel shows the radial action, $\beta_r^{-1}$, as a function of mass (crosses) with dotted lines indicating the scaling~\eqref{eq:J-scale}. The lower panel shows the same for the energy scales. Both panels show good agreement with the expected trends. For clarity we did not over-plot the $\beta_j$ values in Figure \ref{fig:scalings}, but these can be seen to be comparable to $\beta_r$. Because cosmological halos are formed from near-cold collapse, their initial angular momentum will be small. The final dispersion of angular momentum is likely generated through a weak form of the radial orbit instability \citep[e.g.][]{1991MNRAS.248..494S,2006ApJ...653...43M,2008ApJ...685..739B,2009ApJ...704..372B}. Thus the scales of the angular momentum distribution and the radial action distribution are likely to be intimately linked. This is one example of a dynamical consideration which should ultimately be used to link $\vec{\beta}$ values to the initial conditions. \subsection{Real space radial density profiles}\label{sec:real-space} \begin{figure} \includegraphics[width=0.47\textwidth]{rho.pdf} \caption{The real-space density distribution (upper panel) of the 1-component and 2-component maximum entropy solutions (dotted and solid lines respectively) compared to the MW simulation binned density profiles (dots). The softening length in MW is $170\,\mathrm{pc}$, so the profile should be reliable exterior to $\sim 700\,\mathrm{pc}$ \protect\cite[e.g.][]{2003MNRAS.338...14P}. The generic maximum entropy result is a density profile with slowly steepening powerlaw to increasing radii, in agreement with the simulations. The 1-component fit misses the central density cusp, showing that the $\sim 3.5\%$ correction to the low angular momentum orbits (Section \ref{sec:results-j}) is required to reproduce this quintessential feature of simulated dark matter halos. The lower panel shows the cumulative mass as a function of radius.}\label{fig:rho} \end{figure} We have shown that a first-principles maximum entropy argument is capable of describing the phase space distribution of particles in dark matter halos, up to a small correction at low angular momentum. The natural next step is to ask what kind of real-space radial density profiles are implied by this phase space distribution and whether these match the classic rolling-powerlaw shape given by simulations. Calculating the density distribution corresponding to the phase space distribution~\eqref{eq:maxent-solution} is technically involved; a description is given in Appendix \ref{sec:dynam-dens-estim}. There we also explain how the same computer code can be used to calculate equilibrium density profiles from simulations (as opposed to analytic distributions). These profiles are generated subject to our simplifying assumptions of phase mixing and spherical symmetry. They agree well with traditional `binned' estimates of the density, validating the assumptions. Furthermore in the new method, each simulated particle is smeared out over its orbit, resulting in considerably smaller Poisson noise than from traditional binned estimates. Applying the algorithm to the maximum entropy solution, we find that the radial density profile implied by equation \eqref{eq:maxent-solution} follows a shallow power law in the centre and steepens with increasing radius, in qualitative agreement with the behaviour seen in numerical simulations \citep[][and references therein]{1997ApJ...490..493N,2004MNRAS.349.1039N,2009MNRAS.398L..21S}. However, when using the single population phase-space distribution fits, the central slope is too shallow (dotted line, Figure~\ref{fig:rho}). One can obtain higher central densities and inner slopes by changing the parameters, but then the outer slope becomes too steep. On the other hand, if one adopts the incomplete relaxation fit advocated in Section \ref{sec:results-j}, a vastly improved real space density profile is recovered (thick solid line, Figure \ref{fig:rho}). This confirms that the $\sim 3.5\%$ population at low-$j$ is responsible for controlling the cusp. In the discussion below we will recap our current understanding of this issue and give directions for future investigation. \section{Discussion}\label{sec:conclusions} We have shown that maximizing the entropy of a distribution function subject to constraints on total action and energy reproduces the phase space density of particles in simulated dark matter halos. Crucially, there is a clear physical motivation behind this choice of constraints. We started by explaining that, since any equilibrium distribution must be phase-mixed, the late stages of relaxation approach this phase-mixed state. As a consequence $\langle \vec{J} \rangle$ becomes a conserved quantity as equilibrium is approached (Section \ref{sec:conservation-action}). This constitutes a dynamical barrier to continued evolution, preventing energy from being further redistributed. The resulting canonical ensemble (i.e. the maximum entropy solution) is given by equation~\eqref{eq:maxent-solution}. From it we derived two key relationships which can be used to test the phase space of simulated halos, respectively equations~\eqref{eq:MAXENT-Jr-derivative} and~\eqref{eq:MAXENT-angmom}. These were used to demonstrate a close agreement between simulations and theory (Section \ref{sec:results}, \ref{sec:results-j}) over orders of magnitude in probability density, and over a wide range of halo masses from dwarf galaxies to clusters (Section \ref{sec:other-simulations}). We compared to the \cite{1967MNRAS.136..101L} distribution which is obtained when energy can be arbitrarily redistributed between particles, finding that our new canonical ensemble offers a vastly improved fit (see dotted lines in Figure~\ref{fig:Jr}). This strongly suggests that ({\it a}) maximum entropy with suitable dynamical constraints (representing incomplete violent relaxation) is a plausible route to understanding the 6D phase space of dark matter halos; {\it (b)} the newly constrained quantities need not be conserved in general, but must be conserved whenever the system is close to equilibrium, so that their value becomes fixed as the dynamics settle down; and ({\it c}) we have identified a physical argument leading to an important example of these constraints. However we found an overabundance of low angular momentum orbits in the simulations relative to the analytic predictions (Section \ref{sec:results-j}). This implies that there is at least one more important constraint that we have not fully reflected in our analysis. Constructing the radial density profile (Section \ref{sec:real-space}) confirms that, although a small fraction ($\mathrel{\rlap{\lower3pt\hbox{$\sim$} 4\%$) of particles are causing the discrepancy, their existence is essential to understanding the origin of the central density cusps seen in numerical simulations. The correspondingly large density of particles as $j\to 0$ has been found by previous work, notably \cite{2001ApJ...555..240B}, who offered a fitting formula which implies continually increasing particle numbers towards $j=0$. With our higher resolution simulations, we can see that $p_j(j)$ does eventually decrease at sufficiently low $j$, but slower than expected given the shrinking available phase space (Figure \ref{fig:h285-blowup}). Furthermore the large number of particles at low angular momentum can be linked directly to various discrepancies between $\Lambda$CDM theory and observed galaxies \citep{2001MNRAS.326.1205V,2009MNRAS.396..141D}. In particular there must be mechanism to remove the low angular momentum baryons \citep{1986ApJ...303...39D,2001MNRAS.321..471B,2010Natur.463..203G,2010arXiv1010.1004B}. Understanding what causes the accumulation of low angular momentum material in the first place is now added to the list of puzzles in this area. Our maximum entropy picture gives an interesting framework in which to interpret the situation. In Section \ref{sec:results-j} we gave an extensive analysis of equation \eqref{eq:MAXENT-angmom} which suggests two routes to adding material at low $j$. The first option is to appeal to anisotropy (first term on the right hand side); the second is to use a population of particles at low energy (high $\beta_E$ in the second term on the right hand side). We currently prefer the second explanation for the following reason. Particles near the centre have very short orbital periods, which make them decouple from fluctuations on the dynamical timescale of the remainder of the halo. In numerical simulations, the cusps are indeed the first part of the halo to form, and they do not change much at late times \citep{1998ApJ...499L...5M,1998MNRAS.293..337S,2011MNRAS.413.1373W}. \cite{2006MNRAS.368.1931L} construct an explicit 2-phase model of the formation of halos reflecting this differentiation, emphasizing the lack of equilibriation between the inner and outer parts of the halo \cite[see also][]{2011ApJ...743..127L}. Accordingly a timescale constraint could be incorporated from the outset of the maximum entropy argument; we expect this would give similar results to our current approach of fitting a second population. This will be tackled explicitly in future work. The alternative view is that the behaviour at low $j$ may be sensitive to effects of asphericity. This could modify the effective degeneracy. But as we commented in Section \ref{sec:results-j}, the only obvious method available is to pack orbits tightly into a plane, so making the phase space available uniform with $j$, rather than linearly increasing. Numerical results do show halos become more anisotropic towards their centre \citep{2002ApJ...574..538J}. On the other hand, when given a second population to fit, our code does not select this as a viable explanation for the existence of the cusp (Section \ref{sec:results-j}). If a full description of the physics generating low angular momentum orbits can be reached, the work in this paper lays the foundation for a complete description of the collisionless equilibria of dark matter halos. Further questions of interest will include: \begin{itemize} \itemsep 0.2em \item whether and how the constraint vector $\vec{\beta}$ can be derived from initial conditions \citep[which will likely lean heavily on an understanding of the radial orbit instability, e.g.][]{1991MNRAS.248..494S,2006ApJ...653...43M,2008ApJ...685..739B,2009ApJ...704..372B}; \item whether and how maximum entropy arguments can explain the power-law behaviour of the pseudo-phase-space density $\rho(r)/\sigma^3(r)$ \citep{2001ApJ...563..483T}; \item whether and how the phase mixing is maintained to sufficient accuracy to make the $\langle \vec{J} \rangle$ conservation effective over periods of major disturbance \citep[perhaps through chaotic mixing e.g.][]{1996ApJ...471...82M,1997PhRvL..78.3426H}; \item how various moments of the distribution function evolve at higher order in perturbation theory (which is closely related to the previous question); \item how the arguments are changed by adopting an explicitly aspherical background Hamiltonian (assuming this is not already answered when attacking the low-$j$ question); and \item how maximum entropy arguments can be adjoined to microphysical descriptions of cusp destruction \citep{1996MNRAS.283L..72N,2001ApJ...560..636E,2002ApJ...580..627W,2005MNRAS.356..107R,2006Natur.442..539M,2012MNRAS.421.3464P} to shed further light on this essential area of galaxy formation. \end{itemize} Our substantial step forward should give confidence that a full statistical account of the distribution of particles in simulated dark matter halos is achievable without any ad hoc assumptions or modifications to the well-established principle of maximum entropy. Such an account would be extremely powerful for practical and pedagogical aspects of understanding the behaviour of dark matter in the Universe. \section*{Acknowledgements} AP gratefully acknowledges helpful conversations with Steven Gratton, James Binney, Justin Read, Simon White, Carlos Frenk, Julien Devriendt, James Wadsley, Phil Marshall, Julianne Dalcanton, Jorge Pe{\~n}arrubia and John Magorrian, and thanks Kieran Finn for development of computer code for a related project. The MW simulation was run by Alyson Brooks. The GHALO simulation was kindly made available by Joachim Stadel and Doug Potter. FG was funded by NSF grant AST-0908499. NSF grant AST-0607819 and NASA ATP NNX08AG84G. Simulations were run on NASA Advanced Supercomputing facilities. Simulation analysis was performed with the pynbody package ({\tt http://code.google.com/p/pynbody}) on the DiRAC facility jointly funded by STFC, the Large Facilities Capital Fund of BIS and the University of Oxford. This work was supported by the Oxford Martin School and the Beecroft Institute of Particle Astrophysics and Cosmology. \bibliographystyle{mn2e}
\section{Magic State Distillation and Implementation of Rotations} Solovay-Kitaev decomposition \cite{DN05,KitaevEtAl2002,BS12} enables the approximation of any gate using an approximately universal set of elementary gates, e.g., $\{H,T,\Lambda{(X)}\}$, where $\Lambda{(X)}$ denotes the controlled-{\tt NOT} gate. In particular, one can approximate any single-qubit unitary operation using the set $\{H,T\}$. Magic state distillation is then used to produce the magic $\ket H$ states necessary to implement the $T$ gate. We call a state $\ket{\psi}$ \emph{magic} if given $n$ noisy copies of $\ket{\psi}$ and the ability to perform perfect Clifford operations, we can obtain, or ``distill", a purer copy of $\ket{\psi}$ from a Clifford circuit applied to the $n$ noisy copies of $\ket{\psi}$. We can then obtain even purer states which are arbitrarily close to the perfect state by applying the protocol recursively \cite{BK05,MEK05,BH12}. These distilled states can be used to implement non-Clifford operations, e.g., the $T$ gate, as described below. We briefly review how to perform an arbitrary rotation about the $Z$-axis using a resource state, and how to apply the $T$ gate. We assume that Clifford operations are applied perfectly, since they can be implemented fault-tolerantly, and that the resource states are very close to pure. One can use the protocols of \cite{BK05,MEK05} to perform the initial distillation in order to obtain arbitrarily pure resource states for input. We focus on the $+1$ eigenstate of the Hadamard operation, $H$, \begin{eqnarray*} \ket{H} &=& \cos \frac\pi8\ket0 + \sin \frac\pi8 \ket1, \end{eqnarray*} and omit the cost of the initial distillation of the $\ket{H}$ state. We concentrate on single-qubit states found in either the $XZ$- or $XY$-plane of the Bloch sphere. Note that one can easily move a state from one plane to the other by the application of the Clifford $HSHX$ operation. Suppose we start with states $\ket{Z(\theta)}$ and $\ket\psi$: \begin{eqnarray*} \ket{Z(\theta)} &=& \ket0 + \mathrm{e}^{i\theta}\ket1, \\ \ket{\psi} &=& a\ket0 + b\ket1. \end{eqnarray*} The circuit to implement a rotation around the $Z$-axis, presented in Fig.~\ref{fig:RotCircs}, leads to the two-qubit state \begin{eqnarray*} \ket{Z(\theta)}\ket\psi &=& a\ket{00} + b\ket{01} + a\mathrm{e}^{i\theta}\ket{10} + b\mathrm{e}^{i\theta}\ket{11}\\ &\xrightarrow{\Lambda{(X)}}& a\ket{00} + b\ket{11} + a\mathrm{e}^{i\theta}\ket{10} + b\mathrm{e}^{i\theta}\ket{01}. \end{eqnarray*} Upon measurement of the first qubit in the computational basis, we obtain \begin{eqnarray*} & \xrightarrow{m=0} & a\ket{0} + b\mathrm{e}^{i\theta}\ket{1},\\ & \xrightarrow{m=1} & a\mathrm{e}^{i\theta}\ket{0} + b\ket{1} = a\ket{0} + b\mathrm{e}^{-i\theta}\ket{1}, \end{eqnarray*} each with probability $1/2$. Thus, the angle of rotation is chosen at random to be $\theta$ or $-\theta$, up to global phase. An analogous circuit performs a rotation about the $X$-axis. Similar circuits can be found in \cite{F09}. \begin{figure} \[\Qcircuit @C=1em @R=1em { \lstick{\ket{Z(\theta)}} & \gate{X (Z)} & \meter & \rstick{\ket{m}} \cw \\ \lstick{\ket{\psi}} & \ctrl{-1} & \rstick{Z (X) (-1^m \theta)\ket{\psi}} \qw \\ }\] \caption{Circuit randomly implementing a rotation of angle $\pm\theta$ around the $Z(X)$-axis. } \label{fig:RotCircs} \end{figure} As an important example of this procedure, consider the $XY$-plane version of the $\ket H$ state: \begin{equation*} \ket{Z(\pi/4)} = HSHX \ket H = \ket0+\mathrm{e}^{i\pi/4}\ket 1. \end{equation*} Using the circuit in Figure \ref{fig:RotCircs}, we can implement a $Z$-rotation of angle $\pm\pi/4$, producing at random either the $T$ gate or its adjoint, $T^\dagger$. In this particular case, we can deterministically apply the desired gate $T$ or $T^\dagger$ by applying the phase gate $S$, since $ST^\dagger = T$. In general, however, this deterministic correction will not be possible. \section{New states from $\ket H$ states} In this section, we show that we can use a very simple two-qubit Clifford circuit to obtain other non-stabilizer states using only $\ket H$ states as an initial resource, and then show that these states enable the approximation of any single-qubit rotation. We assume that we are provided with perfect copies of $\ket H$. We would like to minimize the number of $\ket H$ states required to implement an arbitrary single-qubit rotation, since these distilled states can be costly to produce. \begin{figure} \ \Qcircuit @C=1em @R=1em { \lstick{\ket{H_0}} & \gate{X} & \meter & \rstick{\ket{0} (\ket{1})} \cw \\ \lstick{\ket{H_i}} & \ctrl{-1} & \rstick{\ket{H_{i+1}} (\ket{H_{i-1}})} \qw \\ }\] \caption{Two-qubit circuit used to obtain new $\ket{H_i}$ states from initial resource states $\ket{H_0}$. Upon measuring the 0 outcome, the output state is $\ket{H_{i+1}}$. Upon measuring the 1 outcome, the output state is $\ket{H_{i-1}}$.}\label{fig:2QbLadderCircs} \end{figure} Consider the circuit of Fig.~\ref{fig:2QbLadderCircs}. One can easily verify that it measures the parity of the two input qubits and decodes the resulting state into the second qubit. We begin by considering the two inputs to be $\ket H$ states. We define $\theta_0=\frac\pi8$ and $\ket H = \ket{H_0} = \cos \theta_0 \ket 0 + \sin \theta_0 \ket 1$. The circuit begins as: \begin{eqnarray*} \ket{H_0}\ket{H_0} &=& \cos^2 \theta_0 \ket{00} + \sin^2 \theta_0 \ket{11}\\ & & + \cos \theta_0 \sin \theta_0 (\ket{01} + \ket{10})\\ & \xrightarrow{\Lambda{(X)}} & \cos^2 \theta_0 \ket{00} + \sin^2 \theta_0 \ket{01}\\ & & + \cos \theta_0 \sin \theta_0 (\ket{11} + \ket{10}). \end{eqnarray*} Upon measurement of the first qubit, we have \begin{eqnarray*} & \xrightarrow{m=0}& \frac{\cos^2\theta_0\ket0 + \sin^2\theta_0\ket1}{\cos^4\theta_0+\sin^4\theta_0},\\ & \xrightarrow{m=1}& \frac{1}{\sqrt2}(\ket0 + \ket1). \end{eqnarray*} We define $\theta_1$ such that \begin{eqnarray*} \cos\theta_1\ket0 + \sin\theta_1\ket1& = & \frac{\cos\theta_0\ket0 + \sin\theta_0\ket1}{\cos^4\theta_0+\sin^4\theta_0}, \end{eqnarray*} from which we deduce $\cot\theta_1=\cot^2\theta_0$. We define $\ket{H_1}=\cos\theta_1\ket0 + \sin\theta_1\ket1$, a non-stabilizer state obtained from $\ket H$ states, Clifford operations, and measurements. If the outcome of the measurement is 1, then we obtain a stabilizer state and discard the output (see Fig.~\ref{fig:2QbLadderCircs}). The two measurement outcomes occur with respective probabilities \begin{eqnarray*} p_0 &=& \cos^4 \theta_0 +\sin^4 \theta_0 =\frac{3}{4},\\ p_1 &=& 1-p_0 = \frac{1}{4}. \end{eqnarray*} We now recurse on this protocol using the non-stabilizer states produced by the previous round of the protocol as part of the input to the circuit of Fig.~\ref{fig:2QbLadderCircs}. We define \begin{eqnarray*} \ket{H_i} &=& \cos \theta_i \ket0 + \sin \theta_i \ket1, \\ \cot \theta_i &=& \cot^{i+1} \theta_0. \end{eqnarray*} If we use as input a copy of $\ket{H_i}$ and a copy of $\ket{H_0}$, we have \begin{eqnarray*} \ket{H_0}\ket{H_i} &=& \cos \theta_0 \cos \theta_i \ket{00} + \sin \theta_0 \sin \theta_i \ket{11}\\ & & + \sin \theta_0 \cos \theta_i \ket{10} +\cos \theta_0 \sin \theta_i \ket{01},\\ & \xrightarrow{\Lambda{(X)}} & \cos \theta_0 \cos \theta_i \ket{00} + \sin \theta_0 \sin \theta_i \ket{01}\\ & & + \sin \theta_0 \cos \theta_i \ket{10} + \cos \theta_0 \sin \theta_i \ket{11}. \end{eqnarray*} Upon measurement of the first qubit, we have \begin{eqnarray*} &\xrightarrow{m=0}& (\cos \theta' \ket0 + \sin \theta' \ket1),\\ &\xrightarrow{m=1}& (\cos \theta'' \ket0 + \sin \theta'' \ket1), \end{eqnarray*} where \begin{eqnarray*} \cot \theta' &=& \cot \theta_i \cot \theta_0 =\cot^{i+2} \theta_0 =\cot \theta_{i+1},\\ \cot \theta'' &=& \cot \theta_i \tan \theta_0 =\cot^{i} \theta_0 =\cot \theta_{i-1}. \end{eqnarray*} Thus, if we measure 0, we obtain the state $\ket{H_{i+1}}$ and if we measure 1, we obtain the state $\ket{H_{i-1}}$. The probability of measuring 0 is given by \begin{eqnarray*} p_{0,i} &=& \cos^2 \theta_i \cos^2 \theta_0 +\sin^2 \theta_i \sin^2 \theta_0. \end{eqnarray*} Note that $\frac{3}{4}\leq p_{0,i} <\cos^2 \frac\pi8=0.853\ldots$. We can view this recursive process as a semi-infinite random walk with biased non-homogeneous probabilities, as Fig.~\ref{fig:RandomWalk} illustrates. Every time a step is taken along this ``ladder" of states, one copy of $\ket H$ is consumed, except at the first node of the ladder (the production of state $\ket{H_0}$) when we require two copies of $\ket H$: if the outcome 1 is measured at the first node, we discard the output and start with two new copies of $\ket H$. \begin{figure} \includegraphics[width=8cm]{RandomWalk}\\ \caption{Process of obtaining other non-stabilizer states from initial $\ket H$ states. A copy of $\ket{H_i}$ and $\ket{H_0}$ probabilistically yield a copy of $\ket{H_{i-1}}$ or $\ket{H_{i+1}}$ using the circuit of Fig.~\ref{fig:2QbLadderCircs}. Each step along the ladder costs one copy of $\ket{H_0}$, except the first one which costs two.}\label{fig:RandomWalk} \end{figure} Table \ref{tab:rots} lists the rotations obtained from the first few $i$ recursions, using the $\ket{H_i}$ states, and Figs.~\ref{fig:AnglesCircle}, \ref{fig:StateLadder} illustrate. Note that there is a factor of two difference between the angle $\theta_i$ involved in the description of the state and the rotation applied, e.g., the $\ket{H_0}$ state is over $\theta_0=\frac\pi8$, and can be used to implement a $\frac\pi4$ rotation. Also, as $0<\theta_i<\frac{\pi}{4}$ $(\forall i)$, the discontinuity of the cotangent is never a problem. \begin{table} \centering \begin{tabular}{|c|l||c|l|} \hline $i$ & $2\theta_i$ & $i$ & $2\theta_i$\\ \hline\hline 0 & $7.853\times10^{-1}$ & 9 & $2.974\times10^{-4}$\\ 1 & $3.398\times10^{-1}$ & 10 & $1.232\times10^{-4}$\\ 2 & $1.419\times10^{-1}$ & 11 & $5.102\times10^{-5}$\\ 3 & $5.886\times10^{-2}$ & 12 & $2.113\times10^{-5}$\\ 4 & $2.439\times10^{-2}$ & 13 & $8.753\times10^{-6}$\\ 5 & $1.010\times10^{-2}$ & 14 & $3.626\times10^{-6}$\\ 6 & $4.184\times10^{-3}$ & 15 & $1.502\times10^{-6}$\\ 7 & $1.733\times10^{-3}$ & 16 & $6.221\times10^{-7}$\\ 8 & $7.179\times10^{-4}$ & 17 & $\dots$\\ \hline \end{tabular} \caption{Rotation by angle $2\theta_i$ implementable using an $\ket{H_i}=\cos\theta_i\ket 0 + \sin\theta_i\ket 1$ state.}\label{tab:rots} \end{table} \begin{figure} \includegraphics[width=6cm]{AnglesCircle}\\ \caption{Red dots: Rotation by angle $2\theta_i$ implementable using an $\ket{H_i}=\cos\theta_i\ket 0 + \sin\theta_i\ket 1$ state.}\label{fig:AnglesCircle} \end{figure} \section{Numerical study of random $Z$-rotations} \label{sec:Cost} In this section, we numerically study the cost of implementing single-qubit rotations. Although the circuit in Fig.~\ref{fig:RotCircs} randomly applies $\theta$ or $-\theta$, we show that our protocol still results in an efficient application of the desired $Z$-rotation. Assume that we have the ideal case, where $\theta_i=2\theta_{i+1}$. In this hypothetical scenario, if we try to apply some rotation using $\ket{H_i}$ and it fails, then we can correct the gate by applying a rotation using $\ket{H_{i-1}}$. If this gate also fails, then we follow with $\ket{H_{i-2}}$, and so on. There are two crucial facts to point out. First, the probability of failing $n$ times in a row scales as $1/2^n$, i.e., it decays exponentially with $n$, such that the expected number of iterations is well-behaved. Second, and very importantly, if a $T$ gate fails (using a $\ket{H_0}$ state), we can always correct deterministically by applying the phase gate $S$ (which is a $Z$-rotation by $\mathrm{e}^{i\pi/2}$). Unfortunately, $\theta_i\neq2\theta_{i+1}$; however, this assumption is not too far from the truth (see Fig.~\ref{fig:StateLadder}). Since we only require that a gate be approximated to a given precision, we can actually apply any $Z$-rotation rapidly and with good accuracy. First, apply the $\ket{H_i}$-rotation that gets you closest to your target angle and then recurse. The next gate to apply will depend on whether the current gate succeeded or not. Due to the property discussed in the previous paragraph, we will rapidly converge towards our target angle. \begin{figure} \includegraphics[width=6 cm]{StateLadder}\\ \caption{Dots: States obtainable by recursively using the circuit of Fig.~\ref{fig:2QbLadderCircs} and only $\ket H$ states as initial input. Full line: exponential decay fit, $\theta_i\sim2.41^{-0.881i}$.}\label{fig:StateLadder} \end{figure} We now present simulation results of obtaining the Z-rotation $Z(\phi)$, where $\phi$ is chosen randomly, in order to characterize the efficiency of the protocol proposed in previous sections. The simulation proceeds as follows: \begin{enumerate} \item Set desired accuracy $\epsilon$. \item Randomly pick a target rotation angle $0<\phi<2\pi$. \item Find the state $\ket{H_i}$ such that $2\theta_i$ is close to $\phi$. \item Simulate an instance of the ladder to obtain that state and add its cost to the offline cost. \item Apply a rotation using the $\ket{H_i}$ state and the circuit of Fig.~\ref{fig:RotCircs} and add one to the online cost. \item Recurse on steps 3 through 5 until the desired accuracy is reached. \end{enumerate} We define the accuracy of the applied rotation $V$ compared to the target rotation $U$ as \begin{eqnarray*} \max_{\ket\psi}D(U\ket{\psi}\bra{\psi}U^\dagger,V\ket{\psi}\bra{\psi}V^\dagger), \end{eqnarray*} where \begin{eqnarray*} D(\rho,\sigma) &=& \frac12 \mathrm{tr}\left(\sqrt{(\rho-\sigma)^\dagger(\rho-\sigma)}\right) \end{eqnarray*} is the trace distance between states $\rho$ and $\sigma$. If $U$ and $V$ are rotations about the same axis, one can show that for small angles of rotation, which will always be our case, this reduces to the difference of rotation angles, $\epsilon=\Delta\phi$. In \cite{BS12}, the distance measure used is \begin{eqnarray*} D(U,V) &=& \sqrt{\frac{2-|\mathrm{tr}(UV^\dagger)|}{2}}. \end{eqnarray*} In the case of rotations about the same axis, it can be reduced to $\sqrt{1-|\cos(\Delta\phi)|}\approx\Delta\phi/\sqrt2$ for small $\Delta\phi$. This conversion between the two measures is important since we later compare performance. To compare the resource cost of our protocol to Solovay-Kitaev decomposition, we define an online and offline cost to apply a unitary gate. The \emph{online cost}, $C_{\textrm{on}}$, is the expected number of non-Clifford gates, or $\ket{H_i}$ states, required to implement the unitary gate on a qubit. The \emph{offline cost}, $C_{\textrm{off}}$, is the total number of distilled $\ket H$ states required to obtain all of the intermediate $\ket{H_i}$ states used to perform the given unitary operation, that is, the sum of the $\ket H$ states used for each ladder process. In our resource costs, we do not include the initial cost to distill $\ket H$ states. For Solovay-Kitaev decomposition, the offline cost is always 0 and the online cost is the total number of $T$ and $T^\dagger$ gates in the decomposition. We ran the simulation for target accuracies ranging between $10^{-12}<\epsilon<10^{-4}$, each time considering a new random angle to produce a sample of $\sim1.8\times10^4$ instances of this protocol. Just like in the case of Solovay-Kitaev decomposition, we suppose that \begin{eqnarray*} C_{\textrm{on}} &\sim& \ln^c(\frac1\epsilon),\\ C_{\textrm{off}} &\sim& \ln^{c'}(\frac1\epsilon), \end{eqnarray*} where $C_{\textrm{on}}$ and $C_{\textrm{off}}$ are the online and offline costs, respectively, such that \begin{eqnarray*} \ln C_{\textrm{on}} &\sim& c\ln\ln(\frac1\epsilon),\\ \ln C_{\textrm{off}} &\sim& c'\ln\ln(\frac1\epsilon). \end{eqnarray*} The results are given in Fig.~\ref{fig:Scaling}. Fits are also presented from which we deduce that $c \sim 1.29$ and $c' \sim 2.27$ for our protocol. \begin{figure} \centering \subfigure[Fit: $\ln(C_\textrm{on})=-0.49 + 1.29\ln(\ln(1/\epsilon))$.]{ \includegraphics[width=6 cm]{ScalingOnline.jpg} \label{fig:ScalingOnline} } \subfigure[Fit: $\ln(C_\textrm{Off})=-0.72 + 2.27\ln(\ln(1/\epsilon))$.]{ \includegraphics[width=6 cm]{ScalingOffline.jpg} \label{fig:ScalingOffline} } \caption{Target accuracies are chosen such that $10^{-12}<\epsilon<10^{-4}$ and the sample size is $\sim 1.8\times10^4$. The clouds of points are used to fit the data according to a linear fit. We obtain $c \sim 1.29$ and $c' \sim 2.27$.}\label{fig:Scaling} \end{figure} As discussed in Section \ref{sec:SK}, both of these scalings represent a significant improvement over the best implementation, to our knowledge, of Solovay-Kitaev decomposition \cite{BS12}, which was itself a significant improvement over the previous implementation of \cite{DN05}. Note that one can implement any single-qubit unitary $U$ using three rotations around the $X$- and $Z$-axes \cite{NC00}: \begin{eqnarray*} U &\propto& X(\alpha)Z(\beta)X(\gamma), \end{eqnarray*} for some angles $\alpha,\beta,\gamma$. We have explicitly shown simulaton results for $Z$-rotations, however $X$-rotations can be obtained at the same cost using the $X$-rotation circuit given in Fig.~\ref{fig:RotCircs}. Thus we can use our protocol to produce each of the three rotations, and produce any desired single-qubit unitary operation. \subsection{Other states} To further reduce the resource costs and their respective scalings, we show that we can use different Clifford circuits to produce new non-Clifford states that can be used as initial resources for the previously presented protocols. We first introduce three new states $\ket{\psi^{0,1,2}_0}$ and discuss how to combine them into the described protocol. Consider the circuit of Fig.~\ref{fig:Psi0Circ}. It is a Clifford circuit to which we input four copies of $\ket H$. The measurement outcome $000$ occurs with probability $3(2+\sqrt2)/32\approx0.320$, otherwise the output is discarded. If the measurement yields result 000, then the produced state is \begin{eqnarray*} \ket{\psi^0_0} &=& \cos\phi^0_0 \ket0+\sin\phi^0_0\ket1,\\ \phi^0_0 &=& \frac\pi2-\cot^{-1}\left(\frac{2+3\sqrt2}{6+5\sqrt2}\right)\approx 0.446. \end{eqnarray*} Since the probability of success is $0.320$ and that every trial consumes four copies of $\ket H$, the average cost to produce $\ket{\psi^0_0}$ is $12.50$ $\ket H$ states. This circuit was designed to measure the stabilizer code presented in Table \ref{tab:Psi0Stab}. Another interesting state can be obtained from the same circuit, substituting one of the input states by a $\ket +$ state as is illustrated by Fig.~\ref{fig:Psi1Circ}. The measurement outcome 000 is obtained with probability $(6+\sqrt2)/32\approx0.232$. The corresponding output state is \begin{eqnarray*} \ket{\psi^1_0} &=& \cos\phi^1_0 \ket0+\sin\phi^1_0 \ket1,\\ \phi^1_0 &=& \frac\pi2-\cot^{-1}\left(\frac{2\sqrt2}{3+\sqrt2}\right)\approx 0.570. \end{eqnarray*} Since the probability of success is $0.232$ and every trial consumes three copies of $\ket H$, the average cost to produce $\ket{\psi^1_0}$ is $12.95$ $\ket H$ states. Fig.~\ref{fig:Psi2Circ} presents another useful circuit. The measurement outcome 000 is obtained with probability $11/32\approx0.344$. The corresponding output state is \begin{eqnarray*} \ket{\psi^2_0} &=& \cos\phi^2_0 \ket0+\sin\phi^2_0 \ket1,\\ \phi^2_0 &=& \frac\pi2-\cot^{-1}\left(\frac{7}{6\sqrt2}\right)\approx 0.690. \end{eqnarray*} The probability of success is $0.344$ and every trial consumes four copies of $\ket H$ such that the average cost to produce $\ket{\psi^2_0}$ is $11.64$ $\ket H$ states. Table \ref{tab:Psi2Stab} presents the stabilizer code in terms of its generators $S$ that are decoded by the circuit. \begin{figure} \[ \Qcircuit @C=1em @R=.5em { \lstick{\ket{H_0}} & \gate{H} & \qw & \gate{X} & \qw & \qw & \ctrl{3} & \meter & \rstick{0} \cw \\ \lstick{\ket{H_0}} & \qw & \ctrl{1} & \qw & \ctrl{2} & \gate{H} & \qw & \meter & \rstick{0} \cw \\ \lstick{\ket{H_0}} & \gate{H} & \gate{X} & \ctrl{-2} & \qw & \ctrl{1} & \qw & \gate{H} & \rstick{H\ket{\psi^0}} \qw \\ \lstick{\ket{H_0}} & \qw & \qw & \qw & \gate{X} & \gate{Z} & \gate{X} & \meter & \rstick{0} \cw \\ }\] \caption{Circuit to produce $\ket{\psi^0_0}$ states. The probability of success is $0.320$ and every trial consumes four copies of $\ket H$ such that the average cost is $12.50$ to produce a copy of $\ket{\psi^0_0}$.}\label{fig:Psi0Circ} \end{figure} \begin{table} \centering \begin{tabular}{|c|c c c c c|} \hline $S$ & $\pm$ & 0 & 1 & 2 & 3\\ \hline\hline $s_0$ & + & X & Z & X & .\\ $s_1$ & + & . & X & Z & X\\ $s_2$ & + & X & . & X & Z\\ $\overline{Z}$ & + & Z & Z & Z & Z\\ \hline \end{tabular} \caption{The stabilizer code decoded by the circuit of Fig.~\ref{fig:Psi0Circ}.}\label{tab:Psi0Stab} \end{table} \begin{figure} \[ \Qcircuit @C=1em @R=.5em { \lstick{\ket{H_0}} & \gate{H} & \qw & \gate{X} & \qw & \qw & \ctrl{3} & \meter & \rstick{0} \cw \\ \lstick{\ket{+}} & \qw & \ctrl{1} & \qw & \ctrl{2} & \gate{H} & \qw & \meter & \rstick{0} \cw \\ \lstick{\ket{H_0}} & \gate{H} & \gate{X} & \ctrl{-2} & \qw & \ctrl{1} & \qw & \gate{H} & \rstick{H\ket{\psi^1}} \qw \\ \lstick{\ket{H_0}} & \qw & \qw & \qw & \gate{X} & \gate{Z} & \gate{X} & \meter & \rstick{0} \cw \\ }\] \caption{Circuit to produce $\ket{\psi^1_0}$ states. The probability of success is $0.232$ and every trial consumes four copies of $\ket H$ such that the average cost is $12.95$ to produce a copy of $\ket{\psi^1_0}$.}\label{fig:Psi1Circ} \end{figure} \begin{figure} \[ \Qcircuit @C=1em @R=.5em { \lstick{\ket{H_0}} & \qw & \ctrl{3} & \ctrl{2} & \gate{H} & \meter & \rstick{0} \cw \\ \lstick{\ket{H_0}} & \gate{X} & \qw & \qw & \gate{X} & \gate{H} & \rstick{\ket{\psi^2}} \qw \\ \lstick{\ket{H_0}} & \ctrl{-1} & \qw & \gate{X} & \qw & \meter & \rstick{0} \cw \\ \lstick{\ket{H_0}} & \qw & \gate{X} & \qw & \ctrl{-2} & \meter & \rstick{0} \cw \\ }\] \caption{Circuit to produce $\ket{\psi^2_0}$ states. The probability of success is $0.344$ and every trial consumes four copies of $\ket H$ such that the average cost is $11.64$ to produce a copy of $\ket{\psi^2_0}$.}\label{fig:Psi2Circ} \end{figure} \begin{table} \centering \begin{tabular}{|c|c c c c c|} \hline $S$ & $\pm$ & 0 & 1 & 2 & 3\\ \hline\hline $s_0$ & + & X & X & X & X\\ $s_1$ & + & Z & . & Z & .\\ $s_2$ & + & Z & . & . & Z\\ $\overline{Z}$ & + & Z & Z & Z & Z\\ \hline \end{tabular} \caption{The stabilizer code decoded by the circuit of Fig.~\ref{fig:Psi2Circ}.}\label{tab:Psi2Stab} \end{table} We will use these states as input states to the circuit given in Fig.~\ref{fig:2QbLadderCircs}, where one of these states is used in place of the $\ket{H_0}$ input state. We start with a copy of $\ket{\psi^i_0}$ and a copy of $\ket{H_0}$. If measurement outcome 1 is obtained, the state is discarded. Otherwise, we obtain \begin{eqnarray*} \ket{\psi^i_1} &=& \cos\phi^i_1 \ket0+\sin\phi^i_1 \ket1,\\ \cot\phi^i_1 &=& \cot\phi^i_0 \cot \theta_0. \end{eqnarray*} Similarly to the $\ket{H_i}$ states, we define \begin{eqnarray*} \ket{\psi^j_i} &=& \cos\phi^j_i \ket0+\sin\phi^j_i \ket1,\\ \cot\phi^j_i &=& \cot\phi^j_0 \cot^i \theta_0. \end{eqnarray*} If we input a copy of $\ket{\psi^j_i}$ and a copy of $\ket{H_0}$, we obtain \begin{eqnarray*} \ket{H_0}\ket{\psi^j_i} &\xrightarrow{\Lambda{(X)}}& \cos \theta_0 \cos \phi^j_i \ket{00}+ \sin \theta_0 \sin \phi^j_i \ket{01}\\ & & + \sin \theta_0 \cos \phi^j_i \ket{10}+ \cos \theta_0 \sin \phi^j_i \ket{11}. \end{eqnarray*} such that the output state obtained is, depending on measurement outcome, \begin{eqnarray*} &\xrightarrow{m=0}& \ket{\psi^j_{i+1}}\\ &\xrightarrow{m=1}& \ket{\psi^j_{i-1}}. \end{eqnarray*} New ``ladders" of states can be obtained using the $\ket{\psi^{0,1,2}_0}$ states as inputs in place of the $\ket{H_0}$ states. Fig.~\ref{fig:StateLadderMulti} shows the four ladders. Table \ref{tab:rotsMulti} lists the rotations obtained from the first few $i$ recursions and Fig.\ref{fig:AnglesCircleMulti} illustrates. We see that the set of possible rotations is more dense. We reproduced the numerical experiment of Section \ref{sec:Cost} with basic offline costs of 12.50, 12.95 and 11.64 for $\ket{\psi^0_0}$, $\ket{\psi^1_0}$, and $\ket{\psi^2_0}$, respectively. The results are presented in Fig.~\ref{fig:ScalingMulti}. Since the set of states is denser, we expected improved scalings for both the online and offline costs. This is indeed the case, we find $c \sim 1.12$ and $c' \sim 1.75$. However, the basic offline costs of our new states $\ket{\psi^i_0}$ are significantly higher; for precision $\sim10^{-4}$, even though the online cost is smaller using the new states, the offline cost is still smaller if we restrict ourselves to the simpler scheme using only $\ket H$ states. For the protocol using the new input states to reduce both the online and offline costs, we need to consider precisions smaller then $\epsilon\approx1.28\times10^{-5}$, see Fig.~\ref{fig:ComparisionHMulti}. \begin{figure} \includegraphics[width=6 cm]{StateLadderMulti}\\ \caption{Dots: States obtainable by recursively using the circuit of Fig.~\ref{fig:2QbLadderCircs} with initial resource states $\ket H$, $\ket{\psi^0}$, $\ket{\psi^1}$ and $\ket{\psi^2}$.}\label{fig:StateLadderMulti} \end{figure} \begin{figure} \includegraphics[width=6cm]{AnglesCircleMulti}\\ \caption{Dots: Rotations implementable using $\ket{H_i}$, $\ket{\psi^0_i}$, $\ket{\psi^1_i}$, $\ket{\psi^2_i}$ states.}\label{fig:AnglesCircleMulti} \end{figure} \begin{table} \centering \begin{tabular}{|c||l|l|l|l|} \hline $i$ & $2\theta_i$ & $2\phi^0_i$ & $2\phi^1_i$ & $2\phi^2_i$\\ \hline\hline 0 & $7.853\times10^{-1}$ & $4.456\times10^{-1}$ & $5.698\times10^{-1}$ & $6.898\times10^{-1}$\\ 1 & $3.398\times10^{-1}$ & $1.871\times10^{-1}$ & $2.415\times10^{-1}$ & $2.954\times10^{-1}$\\ 2 & $1.419\times10^{-1}$ & $7.770\times10^{-2}$ & $1.004\times10^{-1}$ & $1.231\times10^{-1}$\\ 3 & $5.886\times10^{-2}$ & $3.220\times10^{-2}$ & $4.162\times10^{-2}$ & $5.105\times10^{-2}$\\ 4 & $2.439\times10^{-2}$ & $1.334\times10^{-2}$ & $1.724\times10^{-2}$ & $2.115\times10^{-2}$\\ 5 & $1.010\times10^{-2}$ & $5.525\times10^{-3}$ & $7.142\times10^{-3}$ & $8.761\times10^{-3}$\\ 6 & $4.184\times10^{-3}$ & $2.288\times10^{-3}$ & $2.959\times10^{-3}$ & $3.629\times10^{-3}$\\ 7 & $1.733\times10^{-3}$ & $9.479\times10^{-4}$ & $1.225\times10^{-3}$ & $1.503\times10^{-3}$\\ 8 & $7.179\times10^{-4}$ & $3.926\times10^{-4}$ & $5.076\times10^{-4}$ & $6.226\times10^{-4}$\\ \hline \end{tabular} \caption{Rotations implementable using $\ket{H_i},\ket{\psi^0_i},\ket{\psi^1_i},\ket{\psi^2_i}$ states.}\label{tab:rotsMulti} \end{table} \begin{figure} \includegraphics[width=6 cm]{ComparisonHMulti}\\ \caption{Dashed line: Offline (top) and online (bottom) costs for the scheme using only $\ket H$ states. Dotted lines: Offline (top) and online (bottom) costs for the scheme using $\{\ket H,\ket{\psi^0_0},\ket{\psi^1_0},\ket{\psi^2_0}\}$ states. The two top curves cross at $\epsilon\approx1.28\times10^{-5}$.}\label{fig:ComparisionHMulti} \end{figure} \begin{figure} \centering \subfigure[Fit: $\ln(C'_\textrm{on})=-0.78 + 1.12\ln(\ln(1/\epsilon))$.]{ \includegraphics[width=6 cm]{ScalingOnlineMulti.jpg} \label{fig:ScalingOnlineMulti} } \subfigure[Fit: $\ln(C'_\textrm{Off})=0.54 + 1.75\ln(\ln(1/\epsilon))$.]{ \includegraphics[width=6 cm]{ScalingOfflineMulti.jpg} \label{fig:ScalingOfflineMulti} } \caption{Target accuracies are chosen such that $10^{-12}<\epsilon<10^{-4}$ and the sample size is $\sim 1.8\times10^4$. The clouds of points are used to fit the data according to a linear fit. We obtain $c \sim 1.12$ and $c' \sim 1.75$, improving on the results of Fig.\ref{fig:Scaling}.}\label{fig:ScalingMulti} \end{figure} The reason why we do not consider other measurement outcomes in the circuits considered in this section is that in general, potential errors on the $\ket H$ states are amplified by the circuit. Output states in these cases might still prove useful, but a careful analysis of the evolution of errors must be conducted. \subsection{Minimizing the online cost} In this section, we aim to minimize the online cost, at the price of potentially increasing the offline cost. The protocol up to this point can be summarized as follows. Suppose one wants to implement a $Z$-rotation of an arbitrary angle $\phi$ on a logical state $\ket\psi$. One has to implement a sequence of $j$ rotations $\{Z(2\theta_{i_j})\}$ on $\ket\psi$ using the sequence of states $\{\ket{H_{i_j}}\}$, such that $Z(\phi)\approx\prod_j Z(2\theta_{i_j})$. The online cost is given by $\left|\{\ket{H_{i_j}}\}\right|$. Consider instead the following protocol to implement the same rotation by angle $\phi$. Prepare \emph{offline} the state $\ket{Z(\phi)}$ from a copy of $\ket0$. To achieve this, use the protocol described in the previous paragraph to rotate $\ket0$ to $\ket{Z(\phi)}$. Note that you can implement it offline because you are applying rotations on an ancilla state. Then, use $\ket{Z(\phi)}$ \emph{online} to apply the desired rotation. With probability $\frac12$, the rotation $Z(\phi)$ is applied and the online cost is 1. If it fails, correct for it by preparing \emph{offline} the state $\ket{Z(2\phi)}$. Again, with probability $\frac12$, the overall rotation $Z(\phi)$ is applied and the online cost is 2. If it fails, prepare \emph{offline} $\ket{Z(4\phi)}$, and so on. The probability that a number $n$ of iterations is required before success decreases exponentially with $n$. This process is a negative binomial of parameter $p=\frac12$ and the expected number of online rotations before success goes as $\sim\frac1p=2$. \begin{figure} \centering \subfigure[Fit: $\langle C''_\textrm{on}\rangle=1.99$.]{ \includegraphics[width=6 cm]{ScalingOnlineMinOnCost} \label{fig:ScalingOnlineMinOnCost} } \subfigure[Fit: $\ln(C''_\textrm{Off})=1.13 + 1.75\ln(\ln(1/\epsilon))$.]{ \includegraphics[width=6 cm]{ScalingOfflineMinOnCost} \label{fig:ScalingOfflineMinOnCost} } \caption{Target accuracies are chosen such that $10^{-12}<\epsilon<10^{-4}$ and the sample size is $\sim 1.8\times10^3$. (a) The cloud of points is averaged to get the expected number of online rotations. (b) The cloud of point is used to perform a linear fit.}\label{fig:ScalingMinOnCost} \end{figure} We numerically simulated this process for various random angles, $0<\phi<2\pi$, and accuracies, $10^{-12}<\epsilon<10^{-4}$. As Fig.~\ref{fig:ScalingMinOnCost} illustrates, we note that there is no significant change in the online cost for different values of $\epsilon$. Moreover, the expected number of online rotations to apply is roughly two, the result one would expect in this situation. The scaling of the offline cost is the same as that of the previous scheme, as expected. Also, even though the scaling is the same, the actual values of the offline cost are bigger. The shift is $1.13-0.64=0.59$, see Figs. \ref{fig:ScalingOfflineMulti} and \ref{fig:ScalingOfflineMinOnCost}, which corresponds to a factor of $\mathrm{e}^{0.59}\approx1.80$. One would expect a slightly bigger shift of the offline cost by $\ln 2$ since we repeat the scheme twice on average. This suggests that there might exist some favorable correlations between the expected offline cost of an angle $\theta$ and of twice that angle $2\theta$. \section{Erroneous states} In this section, we determine the effect of errors on our resource $\ket H$ states, and their effect on the produced $\ket{H_i}$ states. A priori, the errors might be amplified by the two-qubit circuit of Fig.~\ref{fig:2QbLadderCircs}, however we show this is not the case. The probabilistic and non-homogeneous nature of the presented protocol is not well suited for an analytical study of the evolution of errors on $\ket{H_i}$ states. Instead, we rely on a numerical study for three different types of errors. We use the trace distance on states $\rho$ and $\sigma$, \begin{eqnarray*} D(\rho,\sigma) &=& \frac12 \mathrm{tr}(\sqrt{(\rho-\sigma)^\dagger(\rho-\sigma)}), \end{eqnarray*} to measure the accuracy of the imperfect $\ket{H_i}$ states. We assume Clifford operations are perfect and that errors can only occur on the $\ket H$ states. We consider three types of erroneous states. First, we assume that the mixed state, $\rho^a_0$, is perfectly along the line joining the center of the Bloch sphere and the the perfect state, i.e., \begin{eqnarray*} \rho^a_0(p) &=& (1-p)|H_0\rangle\langle H_0|+p|-H_0\rangle\langle -H_0|, \end{eqnarray*} where $\ket{-H_0}=\sin\frac\pi8 \ket0-\cos\frac\pi8 \ket1$ is the state orthogonal to $\ket{H_0}$. We denote the imperfect version of $\ket{H_i}$ obtained from $\rho^a_0$ states as $\rho^a_i$. If Clifford operations are perfect, we can always bring any mixed state into this form using twirling \cite{MEK05}. However, for the protocol to be of practical interest, we require it to remain stable under the two following types of errors, where we assume that the state is pure, but that the rotation is slightly off of the desired axis by $\delta$: \begin{eqnarray*} \rho^b_0(\delta) &=& \frac12\left(I+\sin\left(\frac\pi4+\delta\right)X+\cos\left(\frac\pi4+\delta\right)Z\right)\\ \rho^c_0(\delta) &=& \frac12\left(I+\sin\frac\pi4 \cos\delta X+\sin\frac\pi4 \sin\delta Y+\cos\frac\pi4 Z\right). \end{eqnarray*} \begin{figure} \includegraphics[width=6 cm]{DistanceConva}\\ \caption{Evolution of the trace distance between imperfect $\rho^a_i$ and perfect $\ket{H_i}$ states. Circles: data for $p=10^{-4}$, $1\leq i\leq 28$. Squares: data for $p=10^{-6}$, $1\leq i\leq 22$. Diamonds: data for $p=10^{-8}$, $1\leq i\leq 16$. The full lines are exponential decay fits: $(2.08*10^{-3})\times2.31^{-i}$ using points $18\leq i\leq 28$, $(1.63*10^{-5})\times2.28^{-i}$ using points $18\leq i\leq 22$ and $(1.26*10^{-7})\times2.24^{-i}$ using points $13\leq i\leq 16$ for the circle, square and diamond data set, respectively. Sample size is $1000$. We conclude that if the initial resource state has desired accuracy, then this is also true of all derived resource states.}\label{fig:DistConva} \end{figure} \begin{figure} \centering \subfigure[Fits: $(1.17*10^{-3})\times2.31^{-i}$, $(1.03*10^{-5})\times2.29^{-i}$ and $(7.50*10^{-8})\times2.25^{-i}$]{ \includegraphics[width=6 cm]{DistanceConvb} \label{fig:DistConvb} } \subfigure[Fits: $(8.28*10^{-4})\times2.31^{-i}$, $(7.32*10^{-6})\times2.30^{-i}$ and $(5.30*10^{-8})\times2.25^{-i}$]{ \includegraphics[width=6 cm]{DistanceConvc} \label{fig:DistConvc} } \caption{Distances between the ideal $\ket{H_i}$ states and the imperfects states $\rho^b_i$ and $\rho^c_i$ respectively. Sample size is $1000$. $\epsilon=10^{-4},10^{-6},10^{-8}$ in both cases.}\label{fig:DistConvbc} \end{figure} We numerically generated pseudo-random instances of the scheme to produce $\ket{H_i}$ states for different values of $i$ and for different noise strengths. We considered 1000 instances for each of the three types of errors and noise strengths $10^{-4},10^{-6}$ and $10^{-8}$. Figures \ref{fig:DistConva} and \ref{fig:DistConvbc} show that the protocol actually reduces the amplitude of possible errors on the resource $\ket H$ states, such that if we start with $\ket H$ meeting our target accuracy, all the subsequent derived $\ket{H_i}$ states will also meet it. This even suggests that for bigger values of $i$, one could use noisier $\ket H$ states and still achieve the desired accuracy. This could make a dramatic difference if it enables one to reduce the number of distillation recursions necessary to prepare the $\ket H$ states. We note a very similar behavior for the three types of errors. The exponential decay of the distance between erroneous and ideal states confirms that the errors are well behaved under the proposed protocol. We note that the bases for the exponential decay of the errors are comparable, but smaller, than the basis for the exponential decay of the angle implemented. So, for a given error rate, there exits a point where the angle of rotation implemented by $\ket{H_i}$ for some $i$ is going to be comparable or smaller to the error on that angle. However, this is not a problem in practice since for, e.g., $\epsilon\sim10^{-4}$, we find $i=150$. For this value of $i$ the angle is $\theta_i\sim10^{-57}$. \section{Comparison to Solovay-Kitaev decomposition} \label{sec:SK} In this section, we compare the performance of the Solovay-Kitaev decomposition of \cite{BS12} and that of the schemes presented in this article. In order to do this, we first consider different $Z$-rotations of angles $\pi/16$, $\pi/128$, and $\pi/1024$ and different accuracies $10^{-4}, 10^{-8}$ and $10^{-12}$. \begin{figure} \subfigure[\newline$\ln(C^Z_\textrm{SK})=-4.88 + 4.41\ln(\ln(1/\epsilon))$\newline$\ln(C^Z_\textrm{On})=-0.49 + 1.29\ln(\ln(1/\epsilon))$\newline$\ln(C^Z_\textrm{Off})=-0.72 + 2.27\ln(\ln(1/\epsilon))$\newline]{ \includegraphics[width=6 cm]{ComparisonSKZ} \label{fig:ComparisonSKZ} } \subfigure[\newline$\ln(C_\textrm{SK})=-2.67 + 3.40\ln(\ln(1/\epsilon))$\newline$\ln(C_\textrm{On})=-0.49 + 1.29\ln(\ln(1/\epsilon))+\ln3$\newline$\ln(C_\textrm{Off})=-0.72 + 2.27\ln(\ln(1/\epsilon))+\ln3$\newline]{ \includegraphics[width=6 cm]{ComparisonSKU} \label{fig:ComparisonSKU} } \caption{Full line: Cost of Solovay Kitaev decompostion (SKD) of (a) random $Z$-rotations and (b) random unitaries as a function of the precision $\epsilon$. Dotted line: Offline costs of (a) random $Z$-rotations and (b) random unitaries. We assume that, loosely speaking, a random unitary is the composition of three random rotations, hence the additional factor of three. Dashed line: Online cost of (a) random $Z$-rotations and (b) random unitaries. (a) For practical values of $\epsilon$, the online cost is significantly smaller than SK. The offline cost is lower than SK when $\epsilon\leq8.71\times10^{-4}$. (b) Again, for practical values of $\epsilon$, the online cost is significantly smaller than SK. The offline cost is lower than SK when $\epsilon\leq2.67\times10^{-7}$.}\label{fig:ComparisonsSK} \end{figure} \begin{figure} \subfigure[\newline$\ln(C'^Z_\textrm{SK})=-4.88 + 4.41\ln(\ln(1/\epsilon))$\newline$\ln(C'^Z_\textrm{On})=-0.78 + 1.12\ln(\ln(1/\epsilon))$\newline$\ln(C'^Z_\textrm{Off})=0.54 + 1.75\ln(\ln(1/\epsilon))$\newline]{ \includegraphics[width=6 cm]{ComparisonSKMultiZ} \label{fig:ComparisonSKMultiZ} } \subfigure[\newline$\ln(C'_\textrm{SK})=-2.67 + 3.40\ln(\ln(1/\epsilon))$\newline$\ln(C'_\textrm{On})=-0.78 + 1.12\ln(\ln(1/\epsilon))+\ln3$\newline$\ln(C'_\textrm{Off})=0.54 + 1.75\ln(\ln(1/\epsilon))+\ln3$\newline]{ \includegraphics[width=6 cm]{ComparisonSKMultiU} \label{fig:ComparisonSKMultiU} } \caption{Full line: Cost of Solovay Kitaev decompostion (SKD) of (a) random $Z$-rotations and (b) random unitaries as a function of the precision $\epsilon$. Dotted line: Offline costs, using $\{\ket H,\ket{\psi^0},\ket{\psi^1},\ket{\psi^2}\}$ as initial resources, of (a) random $Z$-rotations and (b) random unitaries. Dashed line: Online cost, using $\{\ket H,\ket{\psi^0},\ket{\psi^1},\ket{\psi^2}\}$ as initial resources, of (a) random $Z$-rotations and (b) random unitaries. (a) For practical values of $\epsilon$, the online cost is significantly smaller than SK. The offline cost is lower than SK when $\epsilon\leq4.41\times10^{-4}$. (b) Again, for practical values of $\epsilon$, the online cost is significantly smaller than SK. The offline cost is lower than SK when $\epsilon\leq1.03\times10^{-6}$.}\label{fig:ComparisonsSKMulti} \end{figure} \begin{table} \centering \begin{tabular}{|c|c|l|l|l|} \hline $\theta$ & $C$ & $\epsilon=10^{-4}$ & $\epsilon=10^{-8}$ & $\epsilon=10^{-12}$\\ \hline\hline $\pi/16$ & $C_{SK}$ & 43.83 & 2646 & 29120\\ &$C_\textrm{on}$ & 10.20 & 24.52 & 41.95\\ &$C'_\textrm{on}$ & 5.88 & 12.48 & 19.38\\ &$C_\textrm{off}$ & 73.06 & 349.8 & 874.4\\ &$C'_\textrm{off}$ & 98.29 & 306.1 & 595.0\\ \hline\hline $\pi/128$ &$C_{SK}$ & 53.84 & 2879 & 29530\\ &$C_\textrm{on}$ & 5.47 & 18.96 & 39.27\\ &$C'_\textrm{on}$ & 3.32 & 9.27 & 16.91\\ &$C_\textrm{off}$ & 49.18 & 313.0 & 923.9\\ &$C'_\textrm{off}$ & 52.60 & 234.1 & 560.8\\ \hline\hline $\pi/1024$ &$C_\textrm{SK}$ & 128.1 & 2594 & 15075\\ &$C_\textrm{on}$ & 7.99 & 23.08 & 42.93\\ &$C'_\textrm{on}$ & 3.00 & 8.37 & 15.23\\ &$C_\textrm{off}$ & 77.42 & 381.3 & 969.1\\ &$C'_\textrm{off}$ & 65.75 & 245.5 & 530.7\\ \hline \end{tabular} \caption{$C_\textrm{on}$ and $C_\textrm{off}$ are respectively the online and offline costs to implement the $Z$-rotation by angle $\theta$ using only $\ket H$ states, to precision $\epsilon$. $C'_\textrm{on}$ and $C'_\textrm{off}$ refer to the costs for the version of the scheme thats uses $\{\ket H, \ket{\psi^0_0},\ket{\psi^1_0},\ket{\psi^2_0}\}$ as initial resource states. $C_\textrm{SK}$ refers to the extrapolated cost to implement these gates using the results from \cite{BS12}.}\label{tab:examplesRot} \end{table} Table \ref{tab:examplesRot} lists the expected costs. $C_\textrm{on}$ and $C_\textrm{off}$ are respectively the online and offline costs to implement the gate using only $\ket H$ states. $C'_\textrm{on}$ and $C'_\textrm{off}$ refer to the costs for the version of the scheme thats uses $\{\ket H, \ket{\psi^0_0},\ket{\psi^1_0},\ket{\psi^2_0}\}$ as initial resource states. $C_\textrm{SK}$ refers to the extrapolated cost to implement these gates using the results from \cite{BS12}. This extrapolated cost averages over all unitaries ($c\sim3.40$). This is optimistic since the results of \cite{BS12} suggest that $Z$-rotations are actually harder to implement ($c\sim4.3$). Note that in theory, the cost of this algorithm is $O(\log ^c(1/\epsilon)$, where $c=3.97$ \cite{DN05}. In all cases, the online cost is minimal when our proposed scheme enhanced by $\{\ket{\psi^0_0},\ket{\psi^1_0},\ket{\psi^2_0}\}$ is used. For rougher precision, e.g., $10^{-4}$, the offline cost might be such that the total cost is still minimal for the Solovay-Kitaev implementation of \cite{BS12}. For finer precision, e.g., $10^{-8}$ or $10^{-12}$, our proposed protocol becomes very advantageous, since the cost of the Solovay-Kitaev decomposition becomes prohibitive. We also compare the average behavior of the different schemes as Figs.~\ref{fig:ComparisonsSK} and \ref{fig:ComparisonsSKMulti} illustrate. We first start by noting that, loosely speaking, a random unitary is composed of three random rotations, such that the curves presented previously in Fig.~\ref{fig:ComparisionHMulti} must be shifted by $\ln 3$ in general. Fig.~\ref{fig:ComparisonsSK} plots the fit for the Solovay-Kitaev decomposition (solid line), the online cost (dashed) and offline cost (dotted). For all practical accuracies, the online cost of our proposed scheme is consistently the smallest. However, the offline cost becomes advantageous when $\epsilon<8.71\times10^{-4}$ for $Z$-rotations and $\epsilon<2.67\times10^{-7}$ for random unitaries. Fig.~\ref{fig:ComparisonsSKMulti} plots the same for the scheme with additional initial resource states. Similarly, the offline cost becomes advantageous when $\epsilon<4.41\times10^{-4}$ for $Z$-rotations and $\epsilon<1.03\times10^{-6}$ for random unitaries. \section{Conclusion} We have proposed an alternative protocol to Solovay-Kitaev decomposition that results in significantly smaller resource costs, in both the number of required resource states and the depth of the circuit. We have shown a significant improvement on average in the value of $c$, and in many cases the number of distilled states and rotations required to implement a single-qubit unitary gate are reduced. Another advantage of our protocol is that the number of resources required is a ``smoother" function of accuracy, whereas Solovay-Kitaev decomposition is step-like in nature because of the recursion process used in practice. However, note that our protocols and Solovay-Kitaev decomposition are not exclusive. It might be that some unitaries are better implemented using Solovay-Kitaev decomposition, while our scheme is better suited for $Z$-rotations, which occur, among other algorithms, in the quantum Fourier transform. As future research, there are likely a variety of other circuits that enable other ``ladders" of states. One natural extension would be to use the $SH$ eigenstates distilled using the protocols of \cite{BK05,BH12}. Another extension would be to perform a systematic study of ``small" Clifford circuits. Finally, we note that implementing a rotation by choosing the state which results in an angle closest to the target angle is a simple way of achieving our goal, but it is surely suboptimal. An important research direction would be to optimize the sequence of angles required to implement the desired rotation. \section{Acknowledgements} We thank Alex Bocharov for many useful discussions.
\section{Introduction} Fix a rational prime $p\neq 2$ and a PEL datum $\mathcal B=(B,\ast,V,\left(\cdot,\cdot\right),J)$ with auxiliary data $\mathcal B_p=(\mathcal O_B,\mathcal L)$, see Section \ref{SecPelData}. The datum $\mathcal B$ gives rise to a reductive group $G$ over $\QQ$ and a conjugacy class $h$ of homomorphisms $\Res_{\CC/\RR}\GG_m\to G_\RR$. Fix a compact open subgroup $C^p\subset G(\mathbb{A}_f^p)$. From $C^p$ and $\mathcal B_p$ one obtains a compact open subgroup $C\subset G(\mathbb{A}_f)$ and thus a Shimura datum $(G,h,C)$. In \cite[Section 6]{rz}, Rapoport and Zink construct from $\mathcal B,\ \mathcal B_p$ and $C^p$ an integral model $\mathcal A=\mathcal A_{C^p}$ of the Shimura variety associated with $(G,h,C)$. Concretely $\mathcal A$ is defined as a moduli space of abelian schemes with additional structure. In order to study properties of the scheme $\mathcal A$, Rapoport and Zink introduce the so-called local model $M^{\mathrm{loc}}$. It is defined purely in terms of linear algebra and therefore easier to investigate than $\mathcal A$. The schemes $\mathcal A$ and $M^{\mathrm{loc}}$ are related via an intermediate object $\widetilde{\mathcal A}$ fitting into the so called \emph{local model diagram} \begin{equation*} \xymatrix{ &\ar[dl]_{\widetilde{\varphi}}\widetilde{\mathcal A}\ar[dr]^{\widetilde{\psi}}&\\ \mathcal A&&M^{\mathrm{loc}}. } \end{equation*} \'Etale locally on $\mathcal A$, there is a section $s:\mathcal A\to \widetilde{\mathcal A}$ of $\widetilde{\varphi}$ such that the composition $\mathcal A\xrightarrow{s}\widetilde{\mathcal A}\xrightarrow{\widetilde{\psi}} M^{\mathrm{loc}}$ is \'etale. Consequently $\mathcal A$ inherits any property from $M^{\mathrm{loc}}$ which is local for the \'etale topology. In particular, questions about singularities of $\mathcal A$ or the flatness of $\mathcal A$ can equivalently be studied for $M^{\mathrm{loc}}$. Let us mention that recently a purely group-theoretic definition of the local model was given in \cite{pz} by Pappas and Zhu, providing an intriguing new perspective on the local model diagram. The PEL datum also gives rise to an affine smooth group scheme $\Aut(\mathcal L)$, and $\Aut(\mathcal L)$ acts on both $\tilde{A}$ and $M^{\mathrm{loc}}$. The map $\widetilde{\varphi}$ is an $\Aut(\mathcal L)$-torsor, while the map $\widetilde{\psi}$ is $\Aut(\mathcal L)$-equivariant. Denote by $\mathbb{F}$ an algebraic closure of $\mathbb{F}_p$. Via the local model diagram, the decomposition of $M^{\mathrm{loc}}(\mathbb{F})$ into $\Aut(\mathcal L)(\mathbb{F})$-orbits induces the \emph{Kottwitz-Rapoport} (or KR) \emph{stratification} \begin{equation*} \mathcal A(\mathbb{F})=\coprod_{x\in \Aut(\mathcal L)(\mathbb{F})\backslash M^{\mathrm{loc}}(\mathbb{F})}\mathcal A_{x}, \end{equation*} which was first introduced by Ng\^o and Genestier in \cite{genestier_ngo}. The $\Aut(\mathcal L)(\mathbb{F})$-orbits on $M^{\mathrm{loc}}(\mathbb{F})$ admit the following interesting description. In all cases considered thus far (cf.\ the discussion in \cite[\textsection 3.3]{prs}), the special fiber $M^{\mathrm{loc}}_\mathbb{F}$ of $M^{\mathrm{loc}}$ embeds into an \emph{affine flag variety $\mathcal F$}, which is defined as a moduli space of lattice chains over the ring of formal power series $\mathbb{F}[\![u]\!]$. In analogy with the Bruhat decomposition of the classical flag variety, indexed by the finite Weyl group $W$, the affine flag variety admits the \emph{Iwahori decomposition} $\mathcal F(\mathbb{F})=\coprod_{x\in\widetilde{W}}\mathcal C_x$ into Schubert cells $\mathcal C_x$, indexed by the extended affine Weyl group $\widetilde{W}$. It then turns out that $M^{\mathrm{loc}}_\mathbb{F}\subset \mathcal F$ is a disjoint union of Schubert cells and that the decomposition $M^{\mathrm{loc}}(\mathbb{F})=\coprod_{\mathcal C_x\subset M^{\mathrm{loc}}(\mathbb{F})}\mathcal C_x$ coincides with the decomposition of $M^{\mathrm{loc}}(\mathbb{F})$ into $\Aut(\mathcal L)(\mathbb{F})$-orbits. As in the case of the Bruhat decomposition, many properties of the Iwahori decomposition are easily expressed by combinatorial properties of the corresponding index element in $\widetilde{W}$. Notably, the dimension of $\mathcal C_x$ is given by the length $\ell(x)$ of $x$ in $\widetilde{W}$, and the closure relation between Schubert cells is expressed by the Bruhat order on $\widetilde{W}$. We conclude that the same statements hold for the KR stratification on $\mathcal A(\mathbb{F})$. Let us explain in detail one case in which this convenient combinatorial behavior of the KR stratification was fruitfully exploited. For $B=\QQ$ and a complete lattice chain $\mathcal L$, the moduli problem $\mathcal A$ specializes to the Siegel moduli space $\mathcal A_I$ of principally polarized abelian varieties with Iwahori level structure. In \cite{gy2}, Görtz and Yu compute the dimension of the $p$-rank 0 locus in $\mathcal A_I$, and this computation was later generalized in \cite{hamacher} by Hamacher to the case of all $p$-rank strata. The method is the same in both cases: Determine all KR strata contained in a given $p$-rank stratum and compute the maximum of their dimensions. For this method to work one of course needs to know that a $p$-rank stratum is indeed the union of the KR strata contained in it. Thus both papers depend crucially on the result \cite[Th\'eor\`eme 4.1]{genestier_ngo} of Ng\^o and Genestier, which states that indeed the $p$-rank is constant on a KR stratum in $\mathcal A_I$, and also provides an explicit formula for the $p$-rank on a given KR stratum. The subject of this paper is to generalize the result of Ng\^o and Genestier on the relationship between the KR and the $p$-rank stratification to more general PEL data. Let us give an outline of the structure of this paper and of the results that we have obtained. In Sections \ref{SecPelData} through \ref{SecLocalModelGen}, we recall the construction of the local model diagram. We then show the following result. \begin{thm} Let $\mathcal B$ be an arbitrary PEL datum. If $\mathcal L$ is \emph{complete} (in the sense of Definition \ref{DefnCompleteness}), the $p$-rank is constant on a KR stratum. \end{thm} Before being able to state our next result, we need some more notation. Denote by $\mathcal O_K=W(\mathbb{F})$ the Witt ring of $\mathbb{F}$, by $K$ the fraction field of $\mathcal O_K$ and by $\sigma$ the Frobenius on $K$. Denote by $\DD$ the diagonalizable affine group with character group $\QQ$ over $K$. For $b\in G(K)$ denote by $\nu_b:\DD\to G_K$ the corresponding Newton map. By definition, the group $G_K$ acts on $V_K$ and thus $\nu_b$ gives rise to a representation of $\DD$ on $V_K$. Consider the corresponding weight decomposition $V_K=\oplus_{\chi\in\QQ} V_\chi$ and define \begin{equation*} \nu_{b,0}:=\dim_K V_0. \end{equation*} Denote by $I$ the stabilizer of $\mathcal L\otimes\mathcal O_K$ in $G(K)$. For $b\in G(K)$ and $x\in I\backslash G(K)/I$ we denote by $X_x(b)=\{g\in G(K)/I\mathrel{|} g^{-1}b\sigma(g)\in IxI\}$ the affine Deligne-Lusztig variety associated with $b$ and $x$. By interpreting the KR stratification in terms of the relative position of $\mathcal L\otimes\mathcal O_K$ to its image under Frobenius, we show that the KR strata are precisely the fibers of a canonical map $\gamma:\mathcal A(\mathbb{F})\to I\backslash G(K)/I$. Denote by $\mathrm{Perm}\subset I\backslash G(K)/I$ the image of $\gamma$, and for $x\in \mathrm{Perm}$ by $\mathcal A_x=\gamma^{-1}(x)$ the corresponding KR stratum. \begin{thm} Let $x\in\mathrm{Perm}$ and let $b\in G(K)$. Assume that $X_x(b)\neq \emptyset$. Then the $p$-rank on $\mathcal A_{x}$ is constant with value $\nu_{b,0}$. \end{thm} In Sections \ref{SecSym} through \ref{SecUni3} we turn to the aforementioned interpretation of the KR stratification in terms of the affine flag variety. Section \ref{SecSym} deals with the case of the symplectic group. Section \ref{SecUni} (resp.\ \ref{SecUni2}, resp.\ \ref{SecUni3}) deals with the case of a unitary group associated with a ramified (resp.\ inert, resp.\ split) quadratic extension. Let us note that the embedding of $M^{\mathrm{loc}}_\mathbb{F}$ into an affine flag variety has a long history. In particular we want to emphasize that we have greatly profited from the expositions by Pappas and Rapoport in \cite{pr2}, \cite{pr}, \cite{pr3}, and by Smithling in \cite{smithling_unitary_odd}, \cite{smithling_unitary_even}. We have decided to repeat part of their discussions, on the one hand for the convenience of the reader, and on the other to provide several proofs and details that have been omitted in loc.\ cit. Our discussion is quite similar in all cases. We begin with describing in detail the PEL datum at hand, including the Hodge structure and the resulting determinant morphism. We proceed by making explicit the definition of the local model and investigate its base-change to $\mathbb{F}$. We then recall the definition of the affine flag variety in terms of lattice chains and prove in detail that it can also be described as a suitable quotient of loop groups. We conclude the discussion of the local model by embedding it into the affine flag variety and prove that the $\Aut(\mathcal L)$-orbits on $M^{\mathrm{loc}}(\mathbb{F})$ are precisely the Schubert cells contained in $M^{\mathrm{loc}}(\mathbb{F})$. We then prove an explicit formula for the $p$-rank on a KR stratum. Using these explicit formulas and the aforementioned combinatorial structure of the KR stratification, we prove the following geometric results. \subsection{Density of the ordinary locus} It is an interesting question whether the ordinary locus lies dense in $\mathcal A_\mathbb{F}$. In the case of hyperspecial level structure, it has been studied in detail by Wedhorn in \cite{wedhorn}. We focus on the case of Iwahori level structure. In the corresponding Siegel case $\mathcal A_I$, the result \cite[Corollaire 4.3]{genestier_ngo} of Ng\^o and Genestier answers this question affirmatively. On the other hand, Stamm obtains in \cite{stamm} a negative answer in the two-dimensional Hilbert-Blumenthal case. The following result generalizes these two results and explains the general pattern, thereby answering a question by M.~Rapoport. \begin{thm}[Corollary \ref{CorDense}] Assume that $\mathcal B$ is the symplectic PEL datum associated with a totally real extension $F/\QQ$ (see Section \ref{SecPELSym} for a detailed description of $\mathcal B$). Assume (without loss of generality) that there is only a single prime of $\mathcal O_F$ dividing $p$. Then the ordinary locus lies dense in $\mathcal A_\mathbb{F}$ if and only if $p$ is totally ramified in $\mathcal O_F$. \end{thm} \subsection{Dimension of the \texorpdfstring{$p$}{p}-rank 0 locus} As mentioned above, Görtz and Yu use \cite[Th\'eor\`eme 4.1]{genestier_ngo} to compute the dimension of the $p$-rank 0 locus in $\mathcal A_I$, see \cite[Theorem 8.8]{gy2}. By copying their approach and using our formula for the $p$-rank on a KR stratum in the split unitary case, we obtain the following result. \begin{thm}[Theorem \ref{ThmPRank0Locus}] Assume that $\mathcal B$ is the unitary PEL datum of signature $(r,n-r)$ associated with an imaginary quadratic extension of $\QQ$ in which $p$ splits (see Section \ref{SecPELUni3} for a detailed description of $\mathcal B$). Denote by $\mathcal A^{(0)} \subset \mathcal A(\mathbb{F})$ the subset where the $p$-rank of the underlying abelian variety is equal to $0$. Then \begin{equation*} \dim \mathcal A^{(0)}=\min\bigl((r-1)(n-r),r(n-r-1)\bigr). \end{equation*} \end{thm} \subsection{The Hilbert-Blumenthal case} As an illustrative example, we look in Section \ref{SecHilbBlum} at the case of the Hilbert-Blumenthal modular varieties. Without any additional work, we obtain the following result. \begin{thm}[Theorem \ref{TheoremHB}] Let $g\geq 2$ and let $\mathcal A$ be the Hilbert-Blumenthal modular variety with $\Gamma_0(p)$-level structure associated with a totally real extension of degree $g$ of $\QQ$. Denote by $\mathcal A^{(0)} \subset \mathcal A_{\mathbb{F}}$ and $\mathcal A^{(g)} \subset \mathcal A_{\mathbb{F}}$ the subsets where the $p$-rank of the underlying abelian variety is equal to $0$ and $g$, respectively. Then \begin{equation*} \mathcal A_{\mathbb{F}}=\mathcal A^{(0)} \amalg \mathcal A^{(g)}. \end{equation*} The ordinary locus $\mathcal A^{(g)}$ is the union of only two KR strata $\mathcal A_{x_1}$ and $\mathcal A_{x_2}$. Consequently we have \begin{equation*} \mathcal A_{\mathbb{F}}=\overline{\mathcal A}_{x_1}\cup \overline{\mathcal A}_{x_2}\cup \mathcal A^{(0)}. \end{equation*} Here $\overline{\mathcal A}_{x}$ denotes the closure of the KR stratum $\mathcal A_{x}$ in $\mathcal A_{\mathbb{F}}$. Each of $\mathcal A_{\mathbb{F}}, \overline{\mathcal A}_{x_1}, \overline{\mathcal A}_{x_2}$ and $\mathcal A^{(0)}$ is equidimensional of dimension $2^g$. Furthermore, we have \begin{equation*} \overline{\mathcal A}_{x_1}\cap \overline{\mathcal A}_{x_2}\subset \mathcal A^{(0)}. \end{equation*} \end{thm} Taking $g=2$, we recover the result \cite[Theorem 2 (p.\ 408)]{stamm} of Stamm. \subsection*{Acknowledgments} It is my pleasure to express my gratitude to my advisor U.~Görtz. He is the one who introduced me to this area of mathematics, who suggested the topic of this paper and who has provided invaluable support through countless hours of stimulating discussions. I also want to thank T.~Wedhorn for suggesting the point of view taken in Section \ref{SecFormula}, namely to obtain a formula for the $p$-rank on a KR stratum by looking at the Newton stratification, and M.~Rapoport for helpful comments and suggestions. This work was supported by the SFB/TR45 ``Periods, moduli spaces and arithmetic of algebraic varieties'' of the DFG (German Research Foundation). \subsection*{Notation} We fix once and for all a rational prime $p\neq 2$ and an algebraic closure $\mathbb{F}$ of $\mathbb{F}_p$. Let $n\in\NN_{\geq 1}$. \begin{itemize} \item For elements $x_1,\dots,x_n$ of some set and $k_1,\dots k_n\in\NN$, we denote by $(x_1^{(k_1)},\dots,x_n^{(k_n)})$ the tuple \begin{equation*} (\underset{k_1\text{-times}}{\underbrace{x_1,\dots,x_1}}, ,\dots, \underset{k_n\text{-times}}{\underbrace{x_n,\dots,x_n}}). \end{equation*} For a tuple $x\in\ZZ^n$, we denote by $x(i)$ its $i$-th entry. \item For an element $w$ of $S_n$, the symmetric group on $n$ letters, we denote by $A_w=(\delta_{iw(j)})_{ij}$ the corresponding permutation matrix. \item We write \begin{equation*} \quad\widetilde{J}_{2n}= \begin{pmatrix} 0&\widetilde{I_n}\\ -\widetilde{I_n}&0 \end{pmatrix},\quad\text{where}\quad \widetilde{I}_n=\text{anti-diag}(1,\dots,1). \end{equation*} \end{itemize} Let $R$ be a ring and let $R\to R'$ be an $R$-algebra. \begin{itemize} \item We denote the dual of various objects over $R$ by a superscript $\cdot^\vee=\cdot^{\vee,R}$. \item We often denote the base-change from $R$ to $R'$ by a subscript $\cdot_{R'}$. \item If $G$ is a functor on the category of $R'$-algebras, we denote by $\Res_{R'/R} G$ the functor on the category of $R$-algebras with $(\Res_{R'/R} G)(S)=G(S\otimes_R R')$. \item If $F$ is a functor on the category of $R(\!(u)\!)$-algebras (resp.\ $R[\![u]\!]$-algebras), we denote by $\Lf F=\Lf_u F$ (resp.\ $\Lp F=\Lf^+_u F$) the functor on the category of $R$-algebras with $\Lf F(S)=F(S(\!(u)\!))$ (resp.\ $\Lp F(S)=F(S[\![u]\!])$). \item For $\lambda\in\ZZ^n$, we write $u^\lambda=\diag(u^{\lambda(1)}, \dots, u^{\lambda(n)})\in \mathrm{GL}_n(R(\!(u)\!))$. \end{itemize} \section{The general case}\label{SecGeneral} We assume that the reader is familiar with at least the definitions of \cite[3.1-3.27]{rz} and \cite[6.1-6.9]{rz}. The required results on orders in semisimple algebras can all be found in Reiner's excellent \cite{reiner}. In Sections \ref{SecPelData} through \ref{SecLocalModelGen} we recall from \cite{rz} the general setup of integral models of PEL-type Shimura varieties and their local models. \subsection{PEL data}\label{SecPelData} A \emph{PEL datum} consists of the following objects. \begin{enumerate} \item A finite-dimensional semisimple $\QQ$-algebra $B$. \item A positive\footnote{By this we mean that the involution on $B\otimes \RR$ arising from $\ast$ via base-change is a positive involution in the sense of \cite[\textsection 2]{kottwitz_points}.} involution $\ast$ on $B$. \item A finitely generated left $B$-module $V$. We assume that $V\neq 0$. \item A symplectic form $\left(\cdot,\cdot\right):V\times V\to \QQ$ on the underlying $\QQ$-vector space of $V$, such that for all $v,w\in V$ and all $b\in B$ the relation \begin{equation*} \pairt{bv}{w}=\pairt{v}{b^\ast w} \end{equation*} is satisfied. \item An element $J\in \End_{B\otimes \RR}(V\otimes \RR)$ with $J^2+1=0$ such that the bilinear form $\pairt{\cdot}{J\cdot}_\RR:V_\RR\times V_\RR\to \RR$ is symmetric and positive definite. \end{enumerate} We also fix the following data. \begin{enumerate}[label=(\alph*)] \item A $\ZZ$-order $\mathcal O_B$ in $B$ such that $\mathcal O_B\otimes \ZZ_p$ is a maximal $\ZZ_p$-order in $B\otimes\QQ_p$. We assume that $\mathcal O_B\otimes \ZZ_p$ is stable under $\ast$. \item A self-dual multichain $\mathcal L$ of $\mathcal O_B\otimes \ZZ_p$-lattices in $V\otimes\QQ_p$. \end{enumerate} Denote by $G$ the group on the category of $\QQ$-algebras with \begin{equation*} G(R)=\left\{g\in \mathrm{GL}_{B\otimes R}(V\otimes R)\mathrel{|} \exists c=c(g)\in R^\times \left( \begin{aligned} &\forall x,y\in V\otimes R\\ &\pairt{gx}{gy}_R=c\pairt{x}{y}_R \end{aligned} \right) \right\}. \end{equation*} Note that for $g\in G(R)$, the unit $c(g)\in R^\times$ is indeed uniquely determined in view of the assumption $V\neq 0$ and the perfectness of $\left(\cdot,\cdot\right)$, justifying the notation. We also denote by $c:G\to\GG_{m,\QQ}$ the resulting morphism. Let $\Lambda\in \mathcal L$. We deviate slightly from the notation of \cite{rz} in writing $\Lambda^\vee=\{x\in V_{\QQ_p}\mathrel{|} \pairt{x}{\Lambda}_{\QQ_p} \subset\mathcal \ZZ_p\}$ (in loc.\ cit.\ the notation $\Lambda^\ast$ is used instead). We denote by $\left(\cdot,\cdot\right)_{\Lambda}:\Lambda\times \Lambda^\vee\to \ZZ_p$ the restriction of $\left(\cdot,\cdot\right)_{\QQ_p}$. It is a perfect pairing and induces an isomorphism $\Lambda^\vee\xrightarrow{\sim} \Hom_{\ZZ_p}(\Lambda,\ZZ_p)=\Lambda^{\vee,\ZZ_p}$ of $\mathcal O_B\otimes\ZZ_p$-modules, justifying the notation. For $\Lambda\subset \Lambda'$ in $\mathcal L$ we denote by $\rho_{\Lambda',\Lambda}:\Lambda\to \Lambda'$ the inclusion. For $b\in (B\otimes \QQ_p)^\times$ in the normalizer of $\mathcal O_B\otimes\ZZ_p$, denote by $\Lambda^b$ the $\mathcal O_B\otimes\ZZ_p$-module obtained from $\Lambda$ by restriction of scalars with respect to the morphism $\mathcal O_B\otimes\ZZ_p \to \mathcal O_B\otimes\ZZ_p,\ x\mapsto b^{-1}xb$, and let $\vartheta_{\Lambda,b}:\Lambda^b\to b\Lambda$ be the isomorphism given by multiplication with $b$. Then $(\Lambda,\rho_{\Lambda',\Lambda}, \vartheta_{\Lambda,b},\left(\cdot,\cdot\right)_\Lambda)$ is a polarized multichain of $\mathcal O_B\otimes \ZZ_p$-modules of type $(\mathcal L)$ which, by abuse of notation, we also denote by $\mathcal L$. Let $B\otimes\QQ_p=B_1\times \dots\times B_m$ be the decomposition into simple factors. It induces a decomposition \begin{equation}\label{EqOBecomp} \mathcal O_B\otimes\ZZ_p=\mathcal O_{B_1}\times\dots\times\mathcal O_{B_m} \end{equation} and each $O_{B_i}$ is a maximal $\ZZ_p$-order in $B_i$. We also get a decomposition $V\otimes\QQ_p=V_1\times \dots\times V_m$ into left $B_i$-modules $V_i$. Denote by $\mathcal L_i$ the projection of $\mathcal L$ to $V_i$. It is a chain of $\mathcal O_{B_i}$-lattices in $V_i$. For $\Lambda\in \mathcal L$ we denote by $\Lambda=\Lambda_1\times\dots\times\Lambda_m,\ \Lambda_i\in\mathcal L_i$ the corresponding decomposition. Denote by $V_{\CC,\pm i}$ the $(\pm i)$-eigenspace of $J_\CC$. Complex conjugation induces an isomorphism $V_{\CC,i}\to V_{\CC,-i}$ and consequently \begin{equation}\label{EqDimOfVi} \dim_\CC V_{\CC,i}=\dim_\CC V_{\CC,-i}=\frac{1}{2}\dim_\QQ V. \end{equation} Let us quickly recall from \cite[3.23]{rz} the determinant morphism. See \cite[2.3]{diss} for a more detailed discussion. Let $R$ be a ring, $A$ a (not necessarily commutative) $R$-algebra and let $M$ be a left $A$-module which is finite locally free as an $R$-module. Denote by $V=V_A$ the functor on the category of $R$-algebras with $V(S)=A\otimes_R S$. We define a morphism $\det_{M,A}=\det_M:V\to \mathbb{A}^1_R$ on $S$-valued points by \begin{align*} \det_M(S):A\otimes_R S\to S,\quad x\mapsto \det_{S}(M_S\xrightarrow{x \cdot } M_S). \end{align*} For $x\in A$ denote by $\chi_R(x|M)$ the characteristic polynomial of $M\xrightarrow{x \cdot } M$ over $R$. Below we will phrase the determinant condition using characteristic polynomials. This is warranted by the following statement. \begin{prop} Let $A$ be a (not necessarily commutative) $R$-algebra and let $M$ and $N$ be $A$-modules which are finite locally free over $R$. Let $A_0\subset A$ be a generating set of $A$ as an $R$-module. Then $\det_M=\det_N$ if and only if for all $a\in A_0$ we have $\chi_R(a|M)=\chi_R(a|N)$ . \end{prop} \begin{proof} Clear by the existence of Amitsur's formula \cite[Theorem A]{amitsur}, which, in a suitable sense, expresses the characteristic polynomial of a linear combination of endomorphisms in terms of the characteristic polynomials of the summands. \end{proof} As $V_{\CC,-i}$ is a $B\otimes\CC$-module we get a morphism $\det_{V_{\CC,-i}}:V_{B\otimes\CC}\to \mathbb{A}^1_\CC$. Consider the reflex field $E=\QQ(\tr_\CC(b\otimes 1|V_\CC);\;b\in B)$. The morphism $\det_{V_{\CC,-i}}$ is defined over $\mathcal O_E$. Fix a place $\mathcal Q$ of $\mathcal O_E$ lying over $p$. \subsection{Polarized \texorpdfstring{$\mathcal L$}{L}-sets of abelian varieties}\label{SecSelfDual} \begin{defn}\label{DefnLSet} Let $R$ be an $\mathcal O_{E_\mathcal Q}$-algebra. A \emph{polarized $\mathcal L$-set of abelian varieties over $R$} is a pair $(A,\lambda)$, where $A=(A_\Lambda,\varrho_{\Lambda',\Lambda})$ is an $\mathcal L$-set of abelian varieties over $R$ in the sense of \cite[Definition 6.5]{rz}, and where $\lambda:A\to A^\vee$ is a principal polarization in the sense of \cite[Definition 6.6]{rz}. We say that $(A,\lambda)$ is \emph{of determinant $\det_{V_{\CC,-i}}$} if for all $\Lambda\in \mathcal L$ we have an equality \begin{equation*} \det_{\Lie A_\Lambda}=\det_{V_{\CC,-i}}\otimes_{\mathcal O_E} R \end{equation*} of morphisms $V_{\mathcal O_B\otimes R}\to \mathbb{A}^1_R$. We denote by $\mathcal A$ the functor on the category of $\mathcal O_{E_\mathcal Q}$-algebras with $\mathcal A(R)$ the set of isomorphism classes of polarized $\mathcal L$-sets of abelian varieties of determinant $\det_{V_{\CC,-i}}$ over $R$. \end{defn} \begin{remark} After additionally imposing a suitable level structure away from $p$ in the definition of $\mathcal A$, we may (and will) assume that $\mathcal A$ is representable by a quasi-projective scheme over $\mathcal O_{E_\mathcal Q}$, see \cite[Definition 6.9]{rz} and the discussion following it. We have decided not to include this level structure in our notation as it is of no importance for the question of the $p$-rank on a KR stratum. \end{remark} Let $R$ be a ring. For an abelian scheme $A/R$, we denote by $H^{dR}_1(A/R)$ the first de Rham cohomology of $A$. It is part of a canonical short exact sequence \begin{equation}\label{EqHodgedeRham} 0\to \omega_{A^\vee}\to H_1^{dR}(A/R)\to \Lie(A)\to 0, \end{equation} where $\omega_{A^\vee}\subset H_1^{dR}(A/R)$ denotes the Hodge filtration. All terms of \eqref{EqHodgedeRham} are finite locally free $R$-modules. We have $\rk_R H_1^{dR}(A/R)=2\dim_R A$ and $\rk_R \Lie(A)=\rk_R\omega_{A^\vee}=\dim_R A$. \begin{defn}\label{DefnIntermediate} We denote by $\widetilde{\mathcal A}$ the functor on the category of $\mathcal O_{E_\mathcal Q}$-algebras with $\widetilde{\mathcal A}(R)$ the set of isomorphism classes of pairs $(A,\gamma)$, where $A$ is a polarized $\mathcal L$-set of abelian varieties of determinant $\det_{V_{\CC,-i}}$ over $R$ and \begin{equation*} \gamma:H^{dR}_{1}(A)\xrightarrow{\sim} \mathcal L\otimes R \end{equation*} is an isomorphism of polarized multichains of $\mathcal O_B\otimes R$-modules of type $(\mathcal L)$. Denote by $\widetilde{\varphi}:\widetilde{\mathcal A}\to \mathcal A$ the morphism given on $R$-valued points by $\widetilde{\mathcal A}(R)\to \mathcal A(R),\ (A,\gamma)\mapsto A$. \end{defn} $\Aut(\mathcal L)$ acts from the left on $\widetilde{\mathcal A}$ via $g\cdot (A,\gamma)=(A,g\circ \gamma)$ and $\widetilde{\varphi}$ is invariant for this action. \begin{prop}[{\cite[Theorem 2.2]{pappas_arithmetic}}]\label{PropTorsor} The morphism $\widetilde{\varphi}:\widetilde{\mathcal A}\to \mathcal A$ is an $\Aut(\mathcal L)$-torsor for the \'etale topology. In particular $\widetilde{\varphi}(\mathbb{F})$ is an $\Aut(\mathcal L)(\mathbb{F})$-torsor in the set-theoretic sense. \end{prop} \subsection{The local model diagram and the KR stratification}\label{SecLocalModelGen} We will use the following obvious variant of \cite[Definition 3.27]{rz}. \begin{defn}\label{DefnLocalModelGen} The local model $M^{\mathrm{loc}}$ is the functor on the category of $\mathcal O_{E_\mathcal Q}$-algebras with $M^{\mathrm{loc}}(R)$ the set of tuples $(t_\Lambda)_{\Lambda\in\mathcal L}$ of $\mathcal O_B\otimes R$-submodules $t_\Lambda \subset \Lambda_R$ satisfying the following conditions for all $\Lambda \subset \Lambda'$ in $\mathcal L$. \begin{enumerate} \item\label{DefnLocalModelGen-Functoriality} We have $\rho_{\Lambda',\Lambda,R}(t_\Lambda)\subset t_{\Lambda'}$, so that we get a commutative diagram \begin{equation*} \xymatrix{ t_\Lambda\ar[r]\ar[d]&t_{\Lambda'}\ar[d]\\ \Lambda_{R}\ar[r]^{\rho_{\Lambda',\Lambda,R}}&\Lambda'_R. } \end{equation*} \item\label{DefnLocalModelGen-Projectivity} The quotient $\Lambda_{R}/t_\Lambda$ is a finite locally free $R$-module. \item\label{DefnLocalModelGen-Determinant} We have an equality \begin{equation*} \det_{\Lambda_R/t_\Lambda}=\det_{V_{\CC,-i}}\otimes_{\mathcal O_E}R \end{equation*} of morphisms $V_{\mathcal O_{B}\otimes R}\to \mathbb{A}^1_R$. \item\label{DefnLocalModelGen-DualityCondition} Under the pairing $\left(\cdot,\cdot\right)_{\Lambda,R}:\Lambda_{R}\times \Lambda^\vee_{R}\to R$, the submodules $t_\Lambda$ and $t_{\Lambda^\vee}$ pair to zero. \item\label{DefnLocalModelGen-Periodicity} We have $\vartheta_{\Lambda,b,R}(t_\Lambda)= t_{b\Lambda}$ for all $b\in (B\otimes \QQ_p)^\times$ that normalize $\mathcal O_B\otimes\ZZ_p$. \end{enumerate} \end{defn} \begin{remark} We have added the natural condition \ref{DefnLocalModelGen}(\ref{DefnLocalModelGen-Periodicity}), which seems to be missing from \cite[Definition 3.27]{rz}. \end{remark} \begin{remark}\label{RemLocalModelIsScheme} By definition, $M^{\mathrm{loc}}$ is a closed subscheme of a finite product of Grassmannians. In particular $M^{\mathrm{loc}}$ is a projective scheme over $\Spec \mathcal O_{E_\mathcal Q}$. \end{remark} \begin{remark}\label{RemarkDecompLocalModel} Let $R$ be an $\mathcal O_{E_\mathcal Q}$-algebra and $(t_\Lambda)_{\Lambda}\in M^{\mathrm{loc}}(R)$. For $\Lambda\in\mathcal L$ the decomposition $\Lambda=\Lambda_1\times\dots\times \Lambda_m$ induces a decomposition $t_{\Lambda}=t_{\Lambda,1}\times\dots\times t_{\Lambda,m}$ into $\mathcal O_{B_i}\otimes R$-submodules $t_{\Lambda,i}\subset \Lambda_{i,R}$. Let $i\in\{1,\dots,m\}$ and let $\Lambda\subset \Lambda'$ in $\mathcal L$ with $\Lambda_i=\Lambda'_i$. From condition \ref{DefnLocalModelGen}(\ref{DefnLocalModelGen-Functoriality}) we conclude that $t_{\Lambda,i}\subset t_{\Lambda',i}$. From condition \ref{DefnLocalModelGen}(\ref{DefnLocalModelGen-Determinant}) we conclude that $t_{\Lambda,i}$ and $t_{\Lambda',i}$ both have the same rank over $R$. Thus $t_{\Lambda,i}=t_{\Lambda',i}$ in view of \ref{DefnLocalModelGen}(\ref{DefnLocalModelGen-Projectivity}). Consequently we may unambiguously write $t_{\Lambda_i}=t_{\Lambda,i}$. We conclude that the family $(t_{\Lambda})_{\Lambda\in\mathcal L}$ is determined by the tuple of families \begin{equation*} \bigl((t_{\Lambda_1})_{\Lambda_1\in\mathcal L_1},\dots,(t_{\Lambda_m})_{\Lambda_m\in \mathcal L_m}\bigr). \end{equation*} All conditions of Definition \ref{DefnLocalModelGen} with the exception of condition (\ref{DefnLocalModelGen-DualityCondition}) translate into independent conditions on the individual $(t_{\Lambda_i})$. \end{remark} \begin{defn} Denote by $\widetilde{\psi}:\widetilde{\mathcal A}\to M^{\mathrm{loc}}$ the morphism given on $R$-valued points by \begin{align*} \widetilde{\mathcal A}(R)&\to M^{\mathrm{loc}}(R),\\ ((A_\Lambda),(\gamma_\Lambda))&\mapsto (\gamma_\Lambda(\omega_{A_\Lambda^\vee}))_\Lambda. \end{align*} \end{defn} $\Aut(\mathcal L)$ acts from the left on $M^{\mathrm{loc}}$ via $(\varphi_\Lambda)\cdot (t_{\Lambda})=(\varphi_\Lambda(t_\Lambda))$ and $\widetilde{\psi}$ is equivariant for this action. \begin{defn} The diagram \begin{equation*} \xymatrix{ &\ar[dl]_{\widetilde{\varphi}}\widetilde{\mathcal A}\ar[dr]^{\widetilde{\psi}}&\\ \mathcal A&&M^{\mathrm{loc}} } \end{equation*} of $\mathcal O_{E_\mathcal Q}$-schemes is called the \emph{local model diagram}. \end{defn} \begin{remark}[{\cite[Chapter 3]{rz}, cf.\ \cite[Theorem 2.2]{pappas_arithmetic}}]\label{RemLocalModelDiag} The morphisms $\widetilde{\varphi}$ and $\widetilde{\psi}$ are smooth of the same relative dimension. There is, \'etale locally on $\mathcal A$, a section $s:\mathcal A\to \widetilde{\mathcal A}$ of $\widetilde{\varphi}$, such that the composition $\widetilde{\psi}\circ s:\mathcal A\to M^{\mathrm{loc}}$ is \'etale. \end{remark} Consider the decomposition \begin{equation*} M^{\mathrm{loc}}(\mathbb{F})=\coprod_{x\in \Aut(\mathcal L)(\mathbb{F})\backslash M^{\mathrm{loc}}(\mathbb{F})}M^{\mathrm{loc}}_x \end{equation*} into $\Aut(\mathcal L)(\mathbb{F})$-orbits. \begin{remark}\label{RemKRStrataLocallyClosed} Let $x\in \Aut(\mathcal L)(\mathbb{F})\backslash M^{\mathrm{loc}}(\mathbb{F})$. The subset $M^{\mathrm{loc}}_x \subset M^{\mathrm{loc}}(\mathbb{F})$ is locally closed, and we equip it with the reduced scheme structure. By \cite[Theorem 3.16]{rz} the $\mathbb{F}$-group $\Aut(\mathcal L)_\mathbb{F}$ is smooth and affine. Thus $M^{\mathrm{loc}}_x$ is a smooth quasi-projective variety over $\mathbb{F}$. \end{remark} For $x\in \Aut(\mathcal L)(\mathbb{F})\backslash M^{\mathrm{loc}}(\mathbb{F})$, we define $\widetilde{\mathcal A}_x=\widetilde{\psi}(\mathbb{F})^{-1}(M^{\mathrm{loc}}_x)$ and $\mathcal A_x=\widetilde{\varphi}(\mathbb{F})(\widetilde{\mathcal A}_x)$. It follows from Proposition \ref{PropTorsor} that the $\mathcal A_x$ are pairwise disjoint and cover $\mathcal A(\mathbb{F})$ as $x$ runs through $\Aut(\mathcal L)(\mathbb{F})\backslash M^{\mathrm{loc}}(\mathbb{F})$. \begin{defn}\label{DefnKR} The decomposition \begin{equation*} \mathcal A(\mathbb{F})=\coprod_{x\in \Aut(\mathcal L)(\mathbb{F})\backslash M^{\mathrm{loc}}(\mathbb{F})}\mathcal A_x \end{equation*} is called the \emph{Kottwitz-Rapoport} (or KR) \emph{stratification} on $\mathcal A$. \end{defn} \begin{remark}\label{RemarkKRasScheme} By Remark \ref{RemLocalModelDiag} there is, \'etale locally on $\mathcal A_\mathbb{F}$, an \'etale morphism $\beta:\mathcal A_\mathbb{F}\to M^{\mathrm{loc}}_\mathbb{F}$ with $\mathcal A_{x}=\beta^{-1}(M^{\mathrm{loc}}_x)$ for $x\in \Aut(\mathcal L)(\mathbb{F})\backslash M^{\mathrm{loc}}(\mathbb{F})$. Hence the subset $\mathcal A_{x} \subset \mathcal A(\mathbb{F})$ is locally closed, and after equipping it with the reduced scheme structure, $\mathcal A_{x}$ is a smooth variety over $\mathbb{F}$. \end{remark} \subsection{The \texorpdfstring{$p$}{p}-rank on a KR stratum}\label{SecPRankGen} \begin{lem}\label{LemModulesOverSimpleAlgebras} Let $A/\QQ_p$ be a finite simple algebra and let $\mathcal O_A\subset A$ be a maximal $\ZZ_p$-order. Then all simple left $\mathcal O_A$-modules and all simple right $\mathcal O_A$-modules have the same finite cardinality. \end{lem} \begin{proof} By \cite[Theorem 17.3]{reiner} there exist a finite division algebra $D/\QQ_p$, an integer $n\in \NN$ and an isomorphism $A\simeq M^{n\times n}(D)$ inducing an isomorphism $\mathcal O_A\simeq M^{n\times n}(\mathcal O_D)$. Here $\mathcal O_D\subset D$ denotes the unique maximal $\ZZ_p$-order, see \cite[Theorem 12.8]{reiner}. Denote by $\mathfrak p\subset \mathcal O_D$ the unique maximal ideal and by $k=\mathcal O_D/\mathfrak p$ the corresponding residue field, see \cite[Theorem 13.2]{reiner}. By loc.\ cit.\ every simple left (resp.\ right) $\mathcal O_D$-module is isomorphic to $k$. Hence by Morita equivalence (see \cite[\S\S 16]{reiner}) every simple left (resp.\ right) $M^{n\times n}(\mathcal O_D)$-module is isomorphic to $k^n=M^{n\times 1}(k)$ (resp.\ $k^n=M^{1\times n}(k)$). \end{proof} \begin{defn}\label{DefnCompleteness} The multichain $\mathcal L$ is called \emph{complete} if for any two neighbors $\Lambda\subset\Lambda'$ in $\mathcal L$, the quotient $\Lambda'/\Lambda$ is a simple $\mathcal O_B\otimes \ZZ_p$-module. \end{defn} For a finite commutative group scheme $G/\mathbb{F}$, we denote by $G^{e,u}$ the \'etale unipotent and by $G^{i,m}$ the infinitesimal multiplicative part of $G$. Let $\rk_{e,u}(G):=\rk(G^{e,u})$ and $\rk_{i,m}(G):=\rk(G^{i,m})$. \begin{lem}\label{LemKernelVeryBoringGen} Assume that $\mathcal L$ is complete. Let $(A_\Lambda,\varrho_{\Lambda',\Lambda})$ be an $\mathcal L$-set of abelian varieties over $\mathbb{F}$ and let $\Lambda\subset\Lambda'$ be neighbors in $\mathcal L$. Then $K=\ker \varrho_{\Lambda',\Lambda}$ is either \'etale unipotent or infinitesimal multiplicative or infinitesimal unipotent. \end{lem} \begin{proof} The decomposition \eqref{EqOBecomp} induces a decomposition $K=K_1\times\dots\times K_m$ into finite locally free group schemes $K_i$ with actions $\mathcal O_{B_i}\to \End(K_i)$. As $\Lambda$ and $\Lambda'$ are neighbors, there is a unique $i_0\in\{1,\dots,m\}$ with $\Lambda_{i_0}\subsetneq \Lambda'_{i_0}$, and as $\mathcal L$ is complete we know that $\Lambda'_{i_0}/\Lambda_{i_0}$ is a simple left $\mathcal O_{B_{i_0}}$-module. Let $N=|\Lambda'_{i_0}/\Lambda_{i_0}|$. By the definition of an $\mathcal L$-set of abelian varieties we know that $K_i=0$ for $i\neq i_0$ and that $G:=K_{i_0}$ has rank $N$ over $\mathbb{F}$. The action $\mathcal O_{B_{i_0}}\to \End G$ induces on $G(\mathbb{F})$ the structure of a left $O_{B_{i_0}}$-module, and as $|G(\mathbb{F})|\leq \rk G= N$, Lemma \ref{LemModulesOverSimpleAlgebras} implies $|G(\mathbb{F})|\in\{0,N\}$. As $|G(\mathbb{F})|=\rk(G^{e,u})$, we conclude that $G^{e,u}\in \{0,G\}$. Denote by $D(G)$ the Cartier dual of $G$. We also obtain on $D(G)(\mathbb{F})$ the structure of a right $O_{B_{i_0}}$-module and we analogously obtain that $D(G)^{e,u}\in \{0,D(G)\}$. As $D(G)^{e,u}=D(G^{i,m})$, it follows that $G^{i,m}\in\{0,G\}$. \end{proof} \begin{defn}\label{DefnPRank} Let $A/\mathbb{F}$ be an abelian variety. Denote by $[p]_A:A\to A$ the multiplication by $p$ and by $A[p]$ the kernel of $[p]_A$. The integer $\log_p \rk_{e,u}A[p]$ is called the \emph{$p$-rank} of $A$. \end{defn} \begin{prop}\label{ProppRankFirstVersionGen} Assume that $\mathcal L$ is complete. Let $(A_\Lambda,\varrho_{\Lambda',\Lambda})$ be an $\mathcal L$-set of abelian varieties over $\mathbb{F}$. Let $\Lambda\in\mathcal L$ and choose a sequence $p^{-1}\Lambda=\Lambda^{(0)}\supsetneq \Lambda^{(1)}\supsetneq \dots\supsetneq \Lambda^{(k)}=\Lambda$ of neighbors $\Lambda^{(j-1)}\supsetneq \Lambda^{(j)}$ in $\mathcal L$. Define \begin{equation*} J_{e,u}=\{j\in\{1,\dots,k\}\mathrel{|} \ker \varrho_{ \Lambda^{(j-1)}, \Lambda^{(j)}}\text{ is \'etale}\}. \end{equation*} The $p$-rank of $A_\Lambda$ is equal to \begin{equation*} \sum_{j\in J_{e,u}} \log_p |\Lambda^{(j-1)}/\Lambda^{(j)}|. \end{equation*} \end{prop} \begin{proof} By the definition of an $\mathcal L$-set of abelian varieties, there is a periodicity isomorphism $\theta_{p^{-1}\Lambda,p}:A_{p^{-1}\Lambda}\xrightarrow{\sim}A_{\Lambda}$ such that \begin{equation*} [p]_{A_\Lambda}=\theta_{p^{-1}\Lambda,p}\circ \prod_{j=1}^k \varrho_{ \Lambda^{(j-1)}, \Lambda^{(j)}}. \end{equation*} This implies \begin{gather*} \rk_{e,u}A_\Lambda[p]=\prod_{j=1}^k \rk_{e,u}\ker \varrho_{\Lambda^{(j-1)}, \Lambda^{(j)}}. \end{gather*} Lemma \ref{LemKernelVeryBoringGen} and the definition of an $\mathcal L$-set of abelian varieties yield \begin{gather*} \rk_{e,u}\ker \varrho_{ \Lambda^{(j-1)}, \Lambda^{(j)}}= \begin{cases} |\Lambda^{(j-1)}/\Lambda^{(j)}|&\text{if }j\in J_{e,u},\\ 1&\text{otherwise.} \end{cases} \end{gather*} \end{proof} \begin{prop}\label{PropEquivCondGen} Let $A=(A_\Lambda,\varrho_{\Lambda',\Lambda})\in\mathcal A(\mathbb{F})$, choose a lift $\widetilde{A}\in \widetilde{\mathcal A}(\mathbb{F})$ of $A$ under $\widetilde{\varphi}(\mathbb{F})$ and let $(t_\Lambda)=\widetilde{\psi}(\mathbb{F})(\widetilde{A})\in M^{\mathrm{loc}}_x$. Let $\Lambda\subset \Lambda'$ in $\mathcal L$. Then \begin{equation}\label{EqEquivCondGen} \begin{aligned} &\phantom{{}\Leftrightarrow{}}\ker \varrho_{ \Lambda', \Lambda}\text{ is multiplicative}\\ &\Leftrightarrow \rho_{\Lambda',\Lambda,\mathbb{F}}(t_\Lambda)=t_{\Lambda'} \end{aligned} \end{equation} and \begin{equation} \label{EqEquivCondGenII} \begin{aligned} &\phantom{{}\Leftrightarrow{}}\ker \varrho_{ \Lambda', \Lambda}\text{ is \'etale}\\ &\Leftrightarrow \Lambda'_\mathbb{F}=\im \rho_{\Lambda',\Lambda,\mathbb{F}}+t_{\Lambda'}. \end{aligned} \end{equation} \end{prop} \begin{proof} In view of the definition of $\widetilde{\psi}$, the stated equivalences amount to well-known characterizations of the respective conditions on $\ker \varrho_{ \Lambda', \Lambda}$ in terms of the Hodge filtration inside the de Rham cohomology. \end{proof} \begin{cor}\label{CorEquivEtale} Let $x\in \Aut(\mathcal L)(\mathbb{F})\backslash M^{\mathrm{loc}}(\mathbb{F})$ and $(A_\Lambda,\varrho_{\Lambda',\Lambda}),(A'_\Lambda,\varrho'_{\Lambda',\Lambda})\in \mathcal A_x$. Let $\Lambda\subset \Lambda'$ in $\mathcal L$. Then $\ker \varrho_{\Lambda', \Lambda}$ is \'etale if and only if $\ker \varrho'_{\Lambda', \Lambda}$ is \'etale. \end{cor} \begin{proof} For $(t_\Lambda)\in M^{\mathrm{loc}}(\mathbb{F})$, the condition $\Lambda'_\mathbb{F}=\im \rho_{\Lambda',\Lambda,\mathbb{F}}+t_{\Lambda'}$ is clearly invariant under the $\Aut(\mathcal L)(\mathbb{F})$-action on $M^{\mathrm{loc}}(\mathbb{F})$. The claim therefore follows from \eqref{EqEquivCondGenII}. \end{proof} \begin{thm}\label{ThmPRankConstant} Assume that $\mathcal L$ is complete. Let $x\in \Aut(\mathcal L)(\mathbb{F})\backslash M^{\mathrm{loc}}(\mathbb{F})$ and $(A_\Lambda,\varrho_{\Lambda',\Lambda}),(A'_\Lambda,\varrho'_{\Lambda',\Lambda})\in \mathcal A_x$. Let $\Lambda,\Lambda'\in\mathcal L$. Then the $p$-ranks of $A_\Lambda$ and $A'_{\Lambda'}$ coincide. In other words, the $p$-rank is constant on a KR stratum. \end{thm} \begin{proof} The $p$-rank of an abelian variety is an isogeny invariant by \cite[p.\ 147]{mav}, so that it suffices to treat the case $\Lambda=\Lambda'$. The statement then follows from Proposition \ref{ProppRankFirstVersionGen} and Corollary \ref{CorEquivEtale}. \end{proof} \subsection{A formula for the \texorpdfstring{$p$}{p}-rank on a KR stratum}\label{SecFormula} Denote by $K$ the completion of the maximal unramified extension of $\QQ_p$ and by $\mathcal O_K$ the valuation ring of $K$. We identify the residue field of $K$ with $\mathbb{F}$. We denote by $\sigma$ the Frobenius automorphism on $K$, inducing the usual Frobenius $\mathbb{F}\to\mathbb{F},x\mapsto x^p$ on the residue field. By abuse of notation we also denote by $\sigma$ the morphism $G(\sigma):G(K)\to G(K)$. \subsubsection{A \texorpdfstring{$p$}{p}-divisible group analogue} Let $Y/\mathbb{F}$ be a $p$-divisible group. We denote by $\DD(Y)$ the covariant Dieudonn\'e module of $Y$, see for example \cite{demazure}. It is a free $\mathcal O_K$-module, equipped with a $\sigma$-linear endomorphism $F_\DD$ and a $\sigma^{-1}$-linear endomorphism $V_\DD$, satisfying $F_\DD\circ V_\DD=F_\DD\circ V_\DD=p$. Let $Y,Y'$ be $p$-divisible groups over $\mathbb{F}$ and let $\lambda:Y\to (Y')^\vee$ be a morphism. It induces a pairing $\left(\cdot,\cdot\right)_\lambda:\DD(Y)\times\DD(Y')\to \mathcal O_K$ satisfying \begin{equation}\label{EqFVandPair} \pairt{F_\DD x}{y}_\lambda=\sigma\pairt{x}{V_\DD y}_\lambda,\ x\in\DD(Y),y\in\DD(Y'). \end{equation} We denote by $(\overline{\DD}(Y),F_{\overline{\DD}},V_{\overline{\DD}})$ the reduction of $\DD(Y), F_\DD, V_\DD)$ modulo $p$. If $A/\mathbb{F}$ is an abelian variety, let $\DD(A)=\DD(A[p^\infty])$. \begin{prop}[{\cite[Theorem 5.11]{Oda}}]\label{PropOda}\leavevmode \begin{enumerate} \item Let $A/\mathbb{F}$ be an abelian variety. There is a canonical isomorphism \begin{equation}\label{EqOda} \iota=\iota_A:\overline{\DD}(A) \xrightarrow{\sim} H^{dR}_1(A), \end{equation} inducing an isomorphism of short exact sequences \begin{equation}\label{EqDiagOda} \begin{gathered} \xymatrix{ 0\ar[r]& \im V_{\overline{\DD}}\ar[r]\ar[d]^\simeq& \overline{\DD}(A)\ar[r]\ar[d]^\simeq_{\iota}& \Lie(A[p^\infty])\ar[d]^\simeq\ar[r]& 0\\ 0\ar[r]& \omega_{A^\vee}\ar[r]& H_1^{dR}(A)\ar[r]& \Lie(A)\ar[r] &0. } \end{gathered} \end{equation} \item Let $A=(A_\Lambda)\in\mathcal A(\mathbb{F})$. For $\Lambda\in\mathcal L$ let $\iota_\Lambda=\iota_{A_\Lambda}:\overline{\DD}(A_\Lambda)\to H^{dR}_1(A_\Lambda)$ be the isomorphism from \eqref{EqOda}. The resulting morphism \begin{equation}\label{EqOdaChain} \iota=(\iota_\Lambda)_\Lambda:(\overline{\DD}(A_\Lambda))_\Lambda\xrightarrow{\sim} (H^{dR}_{1}(A_\Lambda))_\Lambda \end{equation} is an isomorphism of polarized multichains of $\mathcal O_B\otimes \mathbb{F}$-modules of type $(\mathcal L)$. \end{enumerate} \end{prop} In complete analogy with \cite[Definition 6.5]{rz} and Definition \ref{DefnLSet} we have the notion of a polarized $\mathcal L$-set $Y=(Y_\Lambda, \varrho_{\Lambda',\Lambda},\lambda_\Lambda)$ of $p$-divisible groups of determinant $\det_{V_{\CC,-i}}$ over $\mathbb{F}$. We further obtain a set $\mathcal A_{{\mathrm{p}\text{-}\mathrm{div}}}(\mathbb{F})$ of isomorphism classes of such $Y$. In analogy with Definition \ref{DefnIntermediate}, denote by $\widetilde{\mathcal A}_{\mathrm{p}\text{-}\mathrm{div}}(\mathbb{F})$ the set of isomorphism classes of pairs $(Y,\gamma)$, where $Y$ is as above and $\gamma:\overline{\DD}(Y)\xrightarrow{\sim} \mathcal L\otimes \mathbb{F}$ is an isomorphism of polarized multichains of $\mathcal O_B\otimes \mathbb{F}$-modules of type $(\mathcal L)$. We have the canonical map $\widetilde{\varphi}_{\mathrm{p}\text{-}\mathrm{div}}(\mathbb{F}):\widetilde{\mathcal A}_{\mathrm{p}\text{-}\mathrm{div}}(\mathbb{F})\to \mathcal A_{\mathrm{p}\text{-}\mathrm{div}}(\mathbb{F}),\ (Y,\gamma)\mapsto Y$, and the morphism $\widetilde{\psi}_{\mathrm{p}\text{-}\mathrm{div}}(\mathbb{F}):\widetilde{\mathcal A}_{\mathrm{p}\text{-}\mathrm{div}}(\mathbb{F})\to M^{\mathrm{loc}}(\mathbb{F}),\ (Y,\gamma)\mapsto \gamma(\im V_{\overline{\DD}})$. \begin{equation*} \xymatrix{ &\ar[dl]_{\widetilde{\varphi}_{\mathrm{p}\text{-}\mathrm{div}}}\widetilde{\mathcal A}_{\mathrm{p}\text{-}\mathrm{div}}\ar[dr]^{\widetilde{\psi}_{\mathrm{p}\text{-}\mathrm{div}}}&\\ \mathcal A_{\mathrm{p}\text{-}\mathrm{div}}&&M^{\mathrm{loc}} } \end{equation*} By Proposition \ref{PropOda} we obtain maps $\delta:\mathcal A(\mathbb{F})\to \mathcal A_{\mathrm{p}\text{-}\mathrm{div}}(\mathbb{F}),\ A\mapsto A[p^\infty]$ and $\widetilde{\delta}:\widetilde{\mathcal A}(\mathbb{F})\to \widetilde{\mathcal A}_{\mathrm{p}\text{-}\mathrm{div}}(\mathbb{F}),\ (A,\gamma)\mapsto (A[p^\infty],\gamma\circ \iota)$, and the following diagrams commute. \begin{equation*} \xymatrix{ \widetilde{\mathcal A}(\mathbb{F})\ar[d]_{\widetilde{\varphi}(\mathbb{F})}\ar[r]^-{\widetilde{\delta}}& \widetilde{\mathcal A}_{\mathrm{p}\text{-}\mathrm{div}}(\mathbb{F})\ar[d]^{\widetilde{\varphi}_{\mathrm{p}\text{-}\mathrm{div}}(\mathbb{F})}\\ \mathcal A(\mathbb{F})\ar[r]^-\delta&\mathcal A_{\mathrm{p}\text{-}\mathrm{div}}(\mathbb{F}), }\quad \xymatrix{ \widetilde{\mathcal A}(\mathbb{F})\ar[d]_{\widetilde{\psi}(\mathbb{F})}\ar[r]^-{\widetilde{\delta}}& \widetilde{\mathcal A}_{\mathrm{p}\text{-}\mathrm{div}}(\mathbb{F})\ar[d]^{\widetilde{\psi}_{\mathrm{p}\text{-}\mathrm{div}}(\mathbb{F})}\\ M^{\mathrm{loc}}(\mathbb{F})\ar@{=}[r]&M^{\mathrm{loc}}(\mathbb{F}). } \end{equation*} In absolute analogy with Definition \ref{DefnKR}, we obtain a decomposition \begin{equation*} \mathcal A_{\mathrm{p}\text{-}\mathrm{div}}(\mathbb{F})=\coprod_{x\in \Aut(\mathcal L)(\mathbb{F})\backslash M^{\mathrm{loc}}(\mathbb{F})}\mathcal A_{{\mathrm{p}\text{-}\mathrm{div}},x}, \end{equation*} which we call the \emph{KR stratification} on $\mathcal A_{\mathrm{p}\text{-}\mathrm{div}}(\mathbb{F})$. For $x\in \Aut(\mathcal L)(\mathbb{F})\backslash M^{\mathrm{loc}}(\mathbb{F})$ we have $\mathcal A_x=\delta^{-1}(\mathcal A_{{\mathrm{p}\text{-}\mathrm{div}},x})$. We define as in Definition \ref{DefnPRank} the $p$-rank of a $p$-divisible group over $\mathbb{F}$. The proof of Theorem \ref{ThmPRankConstant} then carries over without any changes to show the following statement. \begin{thm}\label{ThmPRankDivisible} Assume that $\mathcal L$ is complete. Then the $p$-rank is constant on a KR stratum in $\mathcal A_{\mathrm{p}\text{-}\mathrm{div}}(\mathbb{F})$. \end{thm} \subsubsection{The map \texorpdfstring{$\alpha:\mathcal A_{\mathrm{p}\text{-}\mathrm{div}}(\mathbb{F})\to B_I(G)$}{alpha}}\label{SecAlpha} \begin{lem}\label{LemIsomOverK} Let $\mathcal M$ and $\mathcal M'$ be polarized multichains of $\mathcal O_B\otimes\mathcal O_K$-modules of type $(\mathcal L)$. Then the canonical map $\Isom(\mathcal M,\mathcal M')(\mathcal O_K)\to \Isom(\mathcal M,\mathcal M')(\mathbb{F})$ is surjective. In particular $\mathcal M$ and $\mathcal M'$ are isomorphic. \end{lem} \begin{proof} Let $\mathcal I=\Isom(\mathcal M,\mathcal M')$. Clearly $\mathcal I$ is representable by an affine scheme over $\mathcal O_K$. By \cite[Theorem 3.16]{rz} we know that the base-change $\mathcal I\otimes_{\mathcal O_K} \mathcal O_K/p^n$ is in particular formally smooth over $\mathcal O_K/p^n$ for every $n\in\NN$. This easily implies the surjectivity of $\mathcal I(\mathcal O_K)\to \mathcal I(\mathbb{F})$ in view of $\mathcal I(\mathcal O_K)=\lim_{n\in\NN}\mathcal I(\mathcal O_K/p^n)$. The second claim follows, as $\mathcal I(\mathbb{F})\neq \emptyset$ by loc.\ cit. \end{proof} For $g\in G(K)$, we denote by $g\cdot (\mathcal L\otimes\mathcal O_K)$ the tuple $(g(\Lambda\otimes\mathcal O_K))_{\Lambda\in\mathcal L}$ of $\mathcal O_B\otimes\mathcal O_K$-submodules of $V\otimes K$. Let $I=\{g\in G(K)\mathrel{|} g\cdot (\mathcal L\otimes\mathcal O_K)=\mathcal L\otimes\mathcal O_K\}=\{g\in G(K)\mathrel{|} \forall \Lambda\in\mathcal L:\ g(\Lambda\otimes\mathcal O_K)=\Lambda\otimes\mathcal O_K\}$ and $I_0=\{g\in I\mathrel{|} c(g)=1\}$. \begin{lem}\label{LemIvsI0} We have $I=\mathcal O_K^\times I_0$. In particular, any $g\in I$ satisfies $c(g)\in \mathcal O_K^\times$. \end{lem} \begin{proof} Let $g\in I$ and $\Lambda\in \mathcal L$. Then $\Lambda$ is in particular an $\mathcal O_K$-lattice in the $K$-vector space $V_K$ and the fact that $g$ restricts to an automorphism of $\Lambda$ implies that $\det(g)\in \mathcal O_K^\times$. The equation $\det(g)^2=c(g)^{\dim_\QQ V}$ then yields $c(g)\in \mathcal O_K^\times$. As $\mathcal O_K$ is strictly Henselian of residue characteristic different from $2$, there is an $x\in \mathcal O_K^\times$ with $x^2=c(g)$. Then $x^{-1}g\in I_0$, as desired. \end{proof} \begin{lem}\label{LemIvsIprime} Let $g\in I_0$. Then $g$ restricts to an automorphism $g_\Lambda:\Lambda\otimes\mathcal O_K\to \Lambda\otimes\mathcal O_K$ for each $\Lambda\in \mathcal L$. The assignment $g\mapsto (g_\Lambda)_\Lambda$ defines an isomorphism $I_0\xrightarrow{\sim} \Aut(\mathcal L)(\mathcal O_K)$. \end{lem} \begin{proof} If $(\varphi_\Lambda)\in \Aut(\mathcal L)(\mathcal O_K)$, then $\varphi_\Lambda\otimes_{\mathcal O_K}K$ is an element of $G(K)$ which is independent of $\Lambda$. This provides an inverse to the map in question. \end{proof} \begin{prop}[{\cite[3.23]{rz}}]\label{PropDieudonneYieldsPolChainGen} Let $Y=(Y_\Lambda,\varrho_{\Lambda',\Lambda},\lambda_\Lambda)\in\mathcal A_{\mathrm{p}\text{-}\mathrm{div}}(\mathbb{F})$. For $\Lambda\in \mathcal L$ let $\mathcal E_\Lambda:\DD(Y_\Lambda)\times \DD(Y_{\Lambda^\vee})\to \mathcal O_K$ be the pairing induced by $\lambda_\Lambda$. Then $(\DD(Y_\Lambda))_\Lambda$, equipped with the pairings $(\mathcal E_\Lambda)_\Lambda$, is a polarized multichain of $\mathcal O_B\otimes \mathcal O_K$-modules of type $(\mathcal L)$. \end{prop} Define equivalence relations $\sim$ and $\sim_I$ on $G(K)$ by \begin{gather*} x\sim y:\Leftrightarrow \exists g\in G(K):\,y=gx\sigma(g)^{-1},\\ x\sim_I y:\Leftrightarrow \exists i\in I:\,y=ix\sigma(i)^{-1} \end{gather*} and denote by $B(G)=G(K)/\sim$ and $B_I(G)=G(K)/\sim_I$ the corresponding quotients. For an element $b\in G(K)$, we denote by $[b]$ its equivalence class in $B(G)$. Let $(Y_\Lambda)\in \mathcal A_{\mathrm{p}\text{-}\mathrm{div}}(\mathbb{F})$. By Lemma \ref{LemIsomOverK} there is an isomorphism $\varphi=(\varphi_\Lambda)_\Lambda:(\DD(Y_\Lambda))_\Lambda\xrightarrow{\sim}\mathcal L\otimes\mathcal O_K$ of polarized multichains of $\mathcal O_B\otimes \mathcal O_K$-modules of type $(\mathcal L)$. Let $\Lambda\in\mathcal L$. Then $\mathfrak F_\Lambda=\varphi_\Lambda\circ F_\DD\circ \varphi_\Lambda^{-1}$ is a $\sigma$-linear endomorphism of $\Lambda\otimes\mathcal O_K$. By functoriality of the Dieudonn\'e module, the base-change $\mathfrak F_\Lambda\otimes_{\mathcal O_K}K:V\otimes K\to V\otimes K$ is independent of $\Lambda$ and we simply denote it by $\mathfrak F$. In the same way, we obtain from the morphisms $V_\DD$ on the Dieudonn\'e modules a $\sigma^{-1}$-linear endomorphism $\mathfrak V$ of $V\otimes K$ . Let $b=\mathfrak F\circ (\id_V\otimes \sigma)^{-1}$. Then $b$ is a $B\otimes K$-linear endomorphism of $V\otimes K$, and in view of \eqref{EqFVandPair} we have $b\in G(K)$, with $c(b)=p$. If $\varphi':(\DD(Y_\Lambda))_\Lambda\xrightarrow{\sim}\mathcal L\otimes\mathcal O_K$ is another isomorphism, we have $\varphi'=i\circ \varphi$ for some $i\in I_0$, and the resulting element $b'\in G(K)$ will satisfy $b'=ib\sigma(i)^{-1}$. In this way we obtain a well-defined map $\alpha:\mathcal A_{\mathrm{p}\text{-}\mathrm{div}}(\mathbb{F})\to B_I(G)$. The canonical projection $G(K)\to I\backslash G(K)/I$ factors through $B_I(G)$, so that we obtain a map $B_I(G)\xrightarrow{\mathrm{can.}} I\backslash G(K)/I$. Denote by $\gamma$ the composition $\mathcal A_{\mathrm{p}\text{-}\mathrm{div}}(\mathbb{F})\xrightarrow{\alpha}B_I(G)\xrightarrow{\mathrm{can.}} I\backslash G(K)/I$. Denote by $\mathrm{Perm}\subset I\backslash G(K)/I$ the image of $\gamma$. By abuse of notation, we also denote by $\gamma:\mathcal A_{\mathrm{p}\text{-}\mathrm{div}}(\mathbb{F})\to \mathrm{Perm}$ the induced map. Denote by $B_I(G)_\mathrm{Perm} \subset B_I(G)$ the preimage of $\mathrm{Perm}$ under the canonical map $B_I(G)\xrightarrow{\mathrm{can.}} I\backslash G(K)/I$. By abuse of notation, we also denote by $\alpha:\mathcal A_{\mathrm{p}\text{-}\mathrm{div}}(\mathbb{F})\to B_I(G)_\mathrm{Perm}$ the induced map. The situation is visualized by the following commutative diagram, in which the square is cartesian. \begin{equation*} \xymatrix{ & B_I(G)\ar[r]^-{\mathrm{can.}}&I\backslash G(K)/I\\ \mathcal A_{\mathrm{p}\text{-}\mathrm{div}}(\mathbb{F})\ar@/_2pc/[rr]_\gamma\ar[r]^\alpha&B_I(G)_\mathrm{Perm}\ar[r]^-{\mathrm{can.}} \ar@{}[u]|{\bigcup}&\mathrm{Perm}\ar@{}[u]|{\bigcup} } \end{equation*} The following two results show that the map $\alpha:\mathcal A_{\mathrm{p}\text{-}\mathrm{div}}(\mathbb{F})\to B_I(G)_\mathrm{Perm}$ is very close to being a bijection. Our proofs are based on the discussion of the Siegel case by Hoeve in \cite[Chapter 7]{hoeve}. \begin{prop}\label{PropAlphaSurjective} The map $\alpha:\mathcal A_{\mathrm{p}\text{-}\mathrm{div}}(\mathbb{F})\to B_I(G)_\mathrm{Perm}$ is surjective. \end{prop} \begin{proof} Let $\overline{b}\in B_I(G)_\mathrm{Perm}$ and pick any representative $b\in G(K)$ of $\overline{b}$. By \eqref{EqFVandPair} and Lemma \ref{LemIvsI0} there is a unit $v\in \mathcal O_K^\times$ with $c(b)=vp$. Using Lang's Lemma in combination with an approximation argument, we find a $u\in \mathcal O_K^\times$ with $v=(u\sigma(u)^{-1})^2$. After replacing $b$ by $u^{-1}b\sigma(u)$, we may assume that $c(b)=p$. Define $\mathfrak F=b\circ(\id_V\otimes \sigma)$ and $\mathfrak V=p\mathfrak F^{-1}$. Let $\Lambda\in \mathcal L$. Then $\Lambda\otimes\mathcal O_K$ is stable under $\mathfrak F$ and $\mathfrak V$, and we denote by $\mathfrak F_\Lambda$ and $\mathfrak V_\Lambda$ the induced endomorphisms of $\Lambda\otimes\mathcal O_K$. Dieudonn\'e theory implies that the chain $((\Lambda\otimes\mathcal O_K,\mathfrak F_\Lambda, \mathfrak V_\Lambda),\rho_{\Lambda',\Lambda})$ is of the form $\DD(Y)$ for an $\mathcal L$-set $Y=(Y_\Lambda,\varrho_{\Lambda',\Lambda})$ of $p$-divisible groups of determinant $\det_{V_{\CC,-i}}$ over $\mathbb{F}$. From $c(b)=p$, we obtain \begin{equation*} \pairt{\mathfrak F_\Lambda x}{y}_{\Lambda,\mathcal O_K}=\sigma\pairt{x}{\mathfrak V_{\Lambda^\vee} y}_{\Lambda,\mathcal O_K},\ x\in\Lambda\otimes\mathcal O_K,y\in\Lambda^\vee\otimes\mathcal O_K, \end{equation*} compare \eqref{EqFVandPair}. Dieudonn\'e theory then implies that $\left(\cdot,\cdot\right)_{\Lambda,\mathcal O_K}$ is induced by an isomorphism $\lambda_\lambda:Y_\Lambda\to Y_{\Lambda^\vee}^\vee$, and the tuple $\lambda=(\lambda_\Lambda)$ provides us with a polarization of $Y$. The isomorphism class of $(Y,\lambda)$ is the desired preimage of $\overline{b}$ under $\alpha$. \end{proof} \begin{prop} Let $Y=(Y_\Lambda,\varrho_{\Lambda',\Lambda},\lambda_\Lambda)$ and $Y'=(Y'_\Lambda,\varrho'_{\Lambda',\Lambda},\lambda'_\Lambda)$ be two points of $\mathcal A_{\mathrm{p}\text{-}\mathrm{div}}(\mathbb{F})$. Then $\alpha(Y)=\alpha(Y')$ if and only if there exist both an isomorphism $\phi=(\phi_\Lambda):(Y_\Lambda,\varrho_{\Lambda',\Lambda})\to (Y'_\Lambda,\varrho'_{\Lambda',\Lambda})$ of $\mathcal L$-sets of $p$-divisible groups over $\mathbb{F}$ and a unit $u\in\ZZ_p^\times$ such that the following diagram commutes for all $\Lambda\in\mathcal L$. \begin{equation*} \xymatrix{ Y_\Lambda\ar[r]^{\phi_\Lambda}\ar[d]_{u\lambda_\Lambda}&Y'_{\Lambda}\ar[d]^{\lambda'_{\Lambda}}\\ Y_{\Lambda^\vee}^\vee &Y_{\Lambda^\vee}'^\vee \ar[l]_{\phi_{\Lambda^\vee}^\vee}. } \end{equation*} \end{prop} \begin{proof} Choose isomorphisms $\varphi:\DD(Y)\xrightarrow{\sim}\mathcal L\otimes\mathcal O_K$ and $\varphi':\DD(Y')\xrightarrow{\sim}\mathcal L\otimes\mathcal O_K$, and denote by $b$ and $b'$ the resulting elements of $G(K)$ as above. If $\alpha(Y)=\alpha(Y')$, there is an $i\in I$ with $b'=ib\sigma(i)^{-1}$. We have seen above that $c(b)=p=c(b')$, so that $c(i)=\sigma(c(i))$. By Lemma \ref{LemIvsI0} and \cite[Lemma 1.2]{kottwitzI}, the element $u:=c(i)$ lies in $\ZZ_p^\times$. As in Lemma \ref{LemIvsIprime} we consider $i$ as a tuple of endomorphisms $(\Lambda\otimes \mathcal O_K\to \Lambda\otimes \mathcal O_K)_{\Lambda\in \mathcal L}$. The composition $\varphi'^{-1}\circ i\circ \varphi$ corresponds under Dieudonn\'e theory to the desired morphism $\phi$. Similarly for the converse. \end{proof} \subsubsection{The map \texorpdfstring{$\gamma$}{gamma} and the KR stratification} \begin{prop}\label{PropInTermsOfGamma} Two points $Y,Y'\in \mathcal A_{\mathrm{p}\text{-}\mathrm{div}}(\mathbb{F})$ lie in the same KR stratum if and only if $\gamma(A)=\gamma(A')$. \end{prop} \begin{proof} The map $G(K)\to G(K),\ g\mapsto p\sigma^{-1}(g^{-1})$ descends to a well-defined bijection $\tau:I\backslash G(K)/I\to I\backslash G(K)/I$, and it suffices to show the corresponding statement for the composition $\mathcal A_{\mathrm{p}\text{-}\mathrm{div}}(\mathbb{F})\xrightarrow{\gamma} I\backslash G(K)/I\xrightarrow{\tau} I\backslash G(K)/I$ instead of $\gamma$. Let $Y=(Y_\Lambda)_\Lambda\in\mathcal A_{\mathrm{p}\text{-}\mathrm{div}}(\mathbb{F})$. Choose an isomorphism $\varphi:(\DD(Y_\Lambda))_\Lambda\xrightarrow{\sim}\mathcal L\otimes\mathcal O_K$ and denote by $\mathfrak F$ and $\mathfrak V$ the resulting endomorphisms of $V\otimes K$, as above. Let $b\in G(K)$ with $\mathfrak F=b\circ (\id_V\otimes \sigma)$. We have $\mathfrak V=p\sigma^{-1}(b^{-1})\circ (\id_V\otimes\sigma^{-1})$, so that $\mathfrak V(\Lambda\otimes\mathcal O_K)=p\sigma^{-1}(b^{-1})(\Lambda\otimes\mathcal O_K),\ \Lambda\in \mathcal L$. For $\Lambda\in\mathcal L$, denote by $\pi_\Lambda:\Lambda\otimes\mathcal O_K\to \Lambda\otimes\mathbb{F}$ the canonical projection. For a tuple $M=(M_\Lambda)_{\Lambda\in \mathcal L}$ of $\mathcal O_B\otimes\mathcal O_K$-submodules $p\Lambda\otimes\mathcal O_K\subset M_\Lambda\subset \Lambda\otimes\mathcal O_K$ further write $\pi(M)=(\pi_\Lambda(M_\Lambda))_\Lambda$. Note that $M$ and $\pi(M)$ mutually determine each other. By definition the KR stratum that $Y\in \mathcal A_{\mathrm{p}\text{-}\mathrm{div}}(\mathbb{F})$ lies in is given by the $\Aut(\mathcal L)(\mathbb{F})$-orbit of the point $\pi\bigl((\mathfrak V(\Lambda\otimes\mathcal O_K))_\Lambda\bigr)\in M^{\mathrm{loc}}(\mathbb{F})$. By Lemma \ref{LemIsomOverK}, Lemma \ref{LemIvsI0} and Lemma \ref{LemIvsIprime}, this orbit is equal to \begin{equation*} \pi\bigl(\{ip\sigma^{-1}(b^{-1})\cdot(\mathcal L\otimes\mathcal O_K)\mathrel{|} i\in I\}\bigr). \end{equation*} We conclude by noting that for an element $g\in G(K)$, the set $\{ig\cdot(\mathcal L\otimes\mathcal O_K)\mathrel{|} i\in I\}$ and the image of $g$ in $I\backslash G(K)/I$ mutually determine each other. \end{proof} \begin{defn}\label{DefnIndexKR} For $x\in \mathrm{Perm}$, we denote by $\mathcal A_{{\mathrm{p}\text{-}\mathrm{div}},x}=\gamma^{-1}(x)$ and $\mathcal A_{x}=\delta^{-1}(\mathcal A_{{\mathrm{p}\text{-}\mathrm{div}},x})$ the corresponding KR stratum in $\mathcal A_{\mathrm{p}\text{-}\mathrm{div}}(\mathbb{F})$ and $\mathcal A(\mathbb{F})$, respectively. \end{defn} \begin{remark}[{Compare \cite[11.3]{hoeve}}]\label{RemarkVvsF} The normalization of Definition \ref{DefnIndexKR} amounts to indexing the KR stratification by the relative position of $\mathcal L\otimes\mathcal O_K$ to its image under \emph{Frobenius}. This seems to be the natural normalization in the current context, see in particular Remark \ref{RemarkADLvsIntersection} below. In the context of affine flag varieties however, to be explained in the following sections, the natural normalization seems to be by the relative position of $\mathcal L\otimes\mathcal O_K$ to its image under \emph{Verschiebung}, see for example Remark \ref{RemVvsFSym}. This amounts to replacing the map $\gamma$ by the composition $\mathcal A(\mathbb{F})\xrightarrow{\gamma} I\backslash G(K)/I\xrightarrow{g\mapsto p\sigma^{-1}(g^{-1})} I\backslash G(K)/I$, as we have for example done in the proof of Proposition \ref{PropInTermsOfGamma}. Let us note that for the question of the $p$-rank on a KR stratum both normalizations yield the same results. \end{remark} Denote by $r_p:\mathcal A_{\mathrm{p}\text{-}\mathrm{div}}(\mathbb{F})\to\NN$ the map with $r_p((Y_\Lambda)_\Lambda)$ equal to the common $p$-rank of the $Y_\Lambda$. In view Proposition \ref{PropInTermsOfGamma}, Theorem \ref{ThmPRankDivisible} amounts to the following statement. \begin{thm}\label{ThmPRankConstant2} Assume that $\mathcal L$ is complete. The map $r_p:\mathcal A_{\mathrm{p}\text{-}\mathrm{div}}(\mathbb{F})\to \NN$ factors through $\mathcal A_{\mathrm{p}\text{-}\mathrm{div}}(\mathbb{F})\xrightarrow{\gamma}\mathrm{Perm}$. \end{thm} \subsubsection{The Newton stratification} The canonical projection $G(K)\to B(G)$ factors through $B_I(G)$, so that we get a map $B_I(G)\xrightarrow{\mathrm{can.}} B(G)$. Denote by $\beta$ the composition $\mathcal A_{\mathrm{p}\text{-}\mathrm{div}}(\mathbb{F})\xrightarrow{\alpha}B_I(G)\xrightarrow{\mathrm{can.}} B(G)$. \begin{defn} Let $b\in B(G)$. We define $\mathcal N_{{\mathrm{p}\text{-}\mathrm{div}},b}:=\beta^{-1}(b)\subset \mathcal A_{\mathrm{p}\text{-}\mathrm{div}}(\mathbb{F})$ and $\mathcal N_{b}:=(\beta\circ \delta)^{-1}(b)\subset \mathcal A(\mathbb{F})$, and call it the \emph{Newton stratum} associated with $b$ in $\mathcal A_{\mathrm{p}\text{-}\mathrm{div}}(\mathbb{F})$ and $\mathcal A(\mathbb{F})$, respectively. \end{defn} Denote by $\DD$ the diagonalizable affine group with character group $\QQ$ over $K$. Let $b\in G(K)$. We denote by $\nu_b:\DD\to G_K$ the corresponding Newton map, defined in \cite[4.2]{kottwitzI}.\footnote{Note that the discussion in loc.\ cit.\ still remains valid for not necessarily connected reductive groups over $\QQ_p$.} The morphism $\nu_b$ makes $V_K$ into a representation of $\DD$ and we consider the corresponding weight decomposition $V_K=\oplus_{\chi\in\QQ} V_\chi$. We define \begin{equation*} \nu_{b,0}:=\dim_K V_0. \end{equation*} If $g\in G(K)$, we know that $\nu_{gb\sigma(g)^{-1}}=\Int(g)\circ \nu_b$, where $\Int(g):G(K)\to G(K),\ h\mapsto ghg^{-1}$, and consequently $\nu_{b,0}=\nu_{gb\sigma(g)^{-1},0}$. Thus the map $G(K)\to\NN,\ b\mapsto \nu_{b,0}$ factors through $B(G)$, and we also denote by $B(G)\to\NN,\ b\mapsto \nu_{b,0}$ the resulting map. \begin{prop}\label{PropPRankNewton} The map $r_p:\mathcal A_{\mathrm{p}\text{-}\mathrm{div}}(\mathbb{F})\to \NN$ factors as \begin{equation*} \mathcal A_{\mathrm{p}\text{-}\mathrm{div}}(\mathbb{F})\xrightarrow{\beta} B(G) \xrightarrow{b\mapsto \nu_{b,0}}\NN. \end{equation*} In other words, for $b\in B(G)$ the $p$-rank on $\mathcal N_b$ is constant with value $\nu_{b,0}$. \end{prop} \begin{proof} This follows from the fact that the isotypical component of slope 0 in $\DD(G)\otimes_{\mathcal O_K} K$ comes precisely from the \'etale part of $G$, see for instance \cite[IV]{demazure}. \end{proof} \textbf{Assume from now on that $\mathcal L$ is complete.} We can summarize the above discussion in the following commutative diagram, with the dotted arrow coming from Theorem \ref{ThmPRankConstant2}. \begin{equation}\label{EqCommDiagram} \begin{gathered} \xymatrix{ \mathcal A_{\mathrm{p}\text{-}\mathrm{div}}(\mathbb{F})\ar@/^2pc/[rr]^\gamma\ar@/_4pc/[drr]_{r_p}\ar@{>>}[r]^-\alpha\ar[dr]_\beta&B_I(G)_\mathrm{Perm}\ar[r]^-{\mathrm{can.}}\ar[d]_{\mathrm{can.}} &\mathrm{Perm}\ar@{.>}[d]\\ &B(G)\ar[r]^-{b\mapsto \nu_{b,0}}&\NN&\\ } \end{gathered} \end{equation} \begin{defn}\label{DefnAffDL} Let $b\in G(K)$ and $x\in I\backslash G(K)/I$. The \emph{affine Deligne-Lusztig variety associated with $b$ and $x$} is defined by \begin{equation*} X_x(b)=\{g\in G(K)/I\mathrel{|} g^{-1}b\sigma(g)\in IxI\}. \end{equation*} \end{defn} \begin{prop}\label{PropNonemptinessInTermsOfFibers} Let $x\in \mathrm{Perm}$ and $b\in G(K)$. Then the following equivalence holds. \begin{equation*} X_x(b)\neq \emptyset \Leftrightarrow \mathcal A_{{\mathrm{p}\text{-}\mathrm{div}},x}\cap \mathcal N_{{\mathrm{p}\text{-}\mathrm{div}},[b]}\neq\emptyset. \end{equation*} \end{prop} \begin{proof} Follows from the definitions and Proposition \ref{PropAlphaSurjective}. \end{proof} \begin{remark}\label{RemarkADLvsIntersection} Let $x\in \mathrm{Perm}$ and let $b\in G(K)$. Although not established in full generality, it is expected that the following equivalence holds, see \cite[Proposition 12.6]{haines}. \begin{equation}\label{EqEquivNonempty} X_x(b)\neq \emptyset \Leftrightarrow \mathcal A_x\cap \mathcal N_{[b]}\neq \emptyset. \end{equation} The difficulty in proving this equivalence lies in the construction of a suitable $\mathcal L$-set of abelian varieties with prescribed $\mathcal L$-set of $p$-divisible groups. For recent progress in the \emph{unramified} case due to Viehmann and Wedhorn see \cite{vw} \end{remark} \begin{thm}\label{ThmFormulaI} Let $x\in \mathrm{Perm}$ and let $b\in G(K)$. Assume that $X_x(b)\neq \emptyset$. Then the $p$-rank on $\mathcal A_{{\mathrm{p}\text{-}\mathrm{div}},x}$ (and a fortiori on $\mathcal A_x$) is constant with value $\nu_{[b],0}$. \end{thm} \begin{proof} Clear from Proposition \ref{PropNonemptinessInTermsOfFibers}, Theorem \ref{ThmPRankConstant2} and Proposition \ref{PropPRankNewton}. \end{proof} \begin{remark} We expect that $I\subset G(K)$ is an Iwahori subgroup. This is true in the situations to be studied in the following sections, and would provide a more natural view on the set $I\backslash G(K)/I$ occurring above as we could then identify it with a suitable extended affine Weyl group, see \cite[Appendix]{pr}. Proving this statement seems to require a case-by-case analysis. The case of a ramified unitary group has been studied in \cite[\textsection 1.2]{pr3}, and the case of the orthogonal group has been investigated in \cite[\textsection 4.3]{smithling_orthogonal_even}. \end{remark} \subsection{A combinatorial lemma}\label{SecComputing} The following combinatorial result explains the relationship between the abstract formula of Theorem \ref{ThmFormulaI} and the more concrete formulas of the following sections. Let $n\in\NN$ and consider the canonical semidirect product $\widetilde{W}:=S_n\ltimes \ZZ^n$. Let $\Xi$ be a finite cyclic group of order $f$ with generator $\sigma$. We have the shift $\prod_{\xi\in\Xi}\widetilde{W}\to \prod_{\xi\in\Xi}\widetilde{W},\ (x_\xi)_\xi\mapsto (x_{\sigma^{-1}\xi})_\xi$. By abuse of notation, we simply denote it by $\sigma$. \begin{lem}\label{LemCombinatoricsInSemiDirect} Let $(w_\xi)_\xi\in \prod_{\xi\in\Xi} S_n$ and $(\lambda_\xi)_\xi\in \prod_{\xi\in\Xi}\ZZ^n$. Assume that for all $\xi\in \Xi$ and all $1\leq i\leq n$, the following statement holds. \begin{equation}\label{EqInclusionConditionConcrete} \lambda_\xi(i)\geq 0\quad\text{and}\quad (\lambda_\xi(i)=0 \Rightarrow w_\xi(i)\leq i). \end{equation} Let $x=(w_\xi u^{\lambda_\xi})_\xi\in \prod_{\xi\in\Xi}\widetilde{W}$. Choose $N\in\NN_{\geq 1}$ such that $\prod_{k=0}^{Nf-1} \sigma^k(x)\in \prod_{\xi\in\Xi}\ZZ^n$. Consider the element \begin{equation*} \nu=(\nu_\xi)_\xi:=\frac{1}{Nf}\prod_{k=0}^{Nf-1} \sigma^k(x) \end{equation*} of $\prod_{\xi\in\Xi}\QQ_{\geq 0}$. Then for each $1\leq i\leq n$, the following statements are equivalent. \begin{enumerate} \item $\exists \xi\in \Xi:\ \nu_\xi(i)=0$. \item $\forall \xi\in \Xi:\ \nu_\xi(i)=0$. \item $\forall \xi\in \Xi:\ (w_\xi(i)=i\wedge \lambda_\xi(i)=0)$.\qed \end{enumerate} \end{lem} \section{The symplectic case}\label{SecSym} \subsection{Number fields}\label{SecNF} We first fix some notation concerning (extensions of) number fields. Let $K/\QQ$ be a number field. We will always denote by $\mathcal O_{K}$ the ring of integers of $K$. If $\mathcal P$ is a nonzero prime of $\mathcal O_{K}$, we will always denote by $k_{\mathcal P}=\mathcal O_{K}/\mathcal P$ its residue field and by $\rho_{\mathcal P}:\mathcal O_{K}\to k_{\mathcal P}$ the corresponding residue morphism. We further denote by $K_\mathcal P$ the completion of $K$ with respect to $\mathcal P$ and by $\mathcal O_{K_\mathcal P}$ the valuation ring of $K_\mathcal P$. Let $K_0/\QQ$ be a number field and assume that $p\mathcal O_{K_0}=\mathcal P_0^{e_0}$ for a single prime $\mathcal P_0$ of $\mathcal O_{K_0}$ and some $e_0\in\NN$. Denote by $\Sigma_0$ the set of all embeddings $K_0\hookrightarrow \CC$. Fix a finite Galois extension $L/\QQ$ with $K_0\subset L$ and write $G=\Gal(L/\QQ)$ and $H_0=\Gal(L/K_0)$. Fix a prime $\mathcal Q$ of $\mathcal O_L$ lying over $\mathcal P_0$ and denote by $G_\mathcal Q\subset G$ the corresponding decomposition group. \begin{lem}\label{LemResidueMapinNonGaloisCase} There is a unique map $\gamma_0=\gamma_{\mathcal P_0}:\Sigma_0\to \Gal(k_{\mathcal P_0}/\mathbb{F}_p)$ satisfying \begin{equation}\label{EqCharPropofGamma} \forall \sigma\in\Sigma_0\forall a\in\mathcal O_{K_0}:\quad \rho_\mathcal Q(\sigma(a))=\gamma_0(\sigma)(\rho_{\mathcal P_0}(a)). \end{equation} It is surjective and all its fibers have cardinality $e_0$. \end{lem} \begin{proof} Left to the reader. \end{proof} Let $K/K_0$ be a quadratic extension. Denote by $\ast$ the non-trivial element of $\Gal(K/K_0)$. Assume that $\mathcal P_0\mathcal O_{K}=\mathcal P_+\mathcal P_-$ for two distinct primes $\mathcal P_+,\mathcal P_-$ of $\mathcal O_{K}$, say $\mathcal Q\cap \mathcal O_{K}=\mathcal P_+$. Consequently $\mathcal P_-=\mathcal P_+^\ast$. Denote by $\alpha:G\to \Sigma$ the restriction map. Fix a lift $\tau_\ast\in G$ of $\ast$ under $\alpha$. Define subsets $\Sigma_\pm\subset \Sigma$ by $\Sigma_+=\alpha(G_\mathcal QH)$ and $\Sigma_-=\alpha(G_\mathcal Q\tau_\ast H)$. Then $\Sigma=\Sigma_+\amalg \Sigma_-$. We identify $k_{\mathcal P_\pm}$ with $k_{\mathcal P_0}$ via the isomorphism induced by the inclusion $\mathcal O_{K_0}\subset \mathcal O_K$ \begin{lem}\label{LemGammaSplit} There are unique maps $\gamma_\pm:\Sigma_\pm\to \Gal(k_{\mathcal P_0}/\mathbb{F}_p)$ satisfying \begin{equation}\label{EqRedOfSigmaUnify} \forall \sigma\in\Sigma_\pm\forall a\in\mathcal O_K:\ \rho_\mathcal Q(\sigma(a))=\gamma_0(\sigma|_{K_0})(\rho_{\mathcal P_\pm}(a)). \end{equation} \end{lem} \begin{proof} Left to the reader. \end{proof} \subsection{The PEL datum}\label{SecPELSym} Let $g,n\in\NN_{\geq 1}$. We start with the PEL datum consisting of the following objects. \begin{enumerate} \item A totally real field extension $F/\QQ$ of degree $g$. \item The identity involution $\id_F$ on $F$. \item A $2n$-dimensional $F$-vector space $V$. \item The symplectic form $\left(\cdot,\cdot\right):V\times V\to \QQ$ on the underlying $\QQ$-vector space of $V$ constructed as follows: Fix once and for all a symplectic form $\left(\cdot,\cdot\right)':V\times V\to F$ and a basis $\mathfrak E'=(e'_1,\dots,e'_{2n})$ of $V$ such that $\left(\cdot,\cdot\right)'$ is described by the matrix $\widetilde{J}_{2n}$ with respect to $\mathfrak E'$. Define $\left(\cdot,\cdot\right)=\tr_{F/\QQ}\circ\left(\cdot,\cdot\right)'$. \item The $F\otimes\RR$-endomorphism $J$ of $V\otimes\RR$ described by the matrix $-\widetilde{J}_{2n}$ with respect to $\mathfrak E'$. \end{enumerate} \begin{remark} Denote by $\mathrm{GSp}_{\left(\cdot,\cdot\right)'}$ the $F$-group of symplectic similitudes with respect to $\left(\cdot,\cdot\right)'$, and by $c:\mathrm{GSp}_{\left(\cdot,\cdot\right)'}\to \GG_m$ the factor of similitude. Then the reductive $\QQ$-group $G$ associated with the above PEL datum fits into the following cartesian diagram. \begin{equation*} \xymatrix{ G\ar[d]_c \,\ar@{^{(}->}[r] & \Res_{F/\QQ} \mathrm{GSp}_{\left(\cdot,\cdot\right)'}\ar[d]^c\\ \GG_{m,\QQ}\,\ar@{^{(}->}[r]& \Res_{F/\QQ}\GG_{m,F}. } \end{equation*} \end{remark} We assume that $p\mathcal O_F=\mathcal P^e$ for a single prime $\mathcal P$ of $\mathcal O_F$. Denote by $f=[k_\mathcal P:\mathbb{F}_p]$ the corresponding inertia degree, so that $g=ef$. We have $F\otimes\QQ_p=F_\mc P$ and $\mathcal O_F\otimes\ZZ_p=\mathcal O_{F_\mathcal P}$. Fix once and for all a uniformizer $\pi$ of $\mathcal O_{F}\otimes \ZZ_{(p)}$. Denote by $\mathfrak C=\mathfrak C_{\mathcal O_{F_\mathcal P}\mathrel{|} \ZZ_p}$ the inverse different of the extension $F_\mc P/\QQ_p$. Fix a generator $\delta$ of $\mathfrak C$ over $\mathcal O_{F_\mathcal P}$ and define a basis $(e_1,\ldots,e_{2n})$ of $V_{\QQ_p}$ over $F_\mc P$ by $e_i=e'_i,\ e_{n+i}=\delta e'_{n+i},\ 1\leq i\leq n$. Let $0\leq i< 2n$. We denote by $\Lambda_i$ the $\mathcal O_{F_\mathcal P}$-lattice in $V_{\QQ_p}$ with basis \begin{equation*} \mathfrak E_i=(\pi^{-1}e_1,\ldots,\pi^{-1}e_i,e_{i+1},\ldots,e_{2n}). \end{equation*} For $k\in\ZZ$ we further define $\Lambda_{2nk+i}=\pi^{-k}\Lambda_i$ and we denote by $\mathfrak E_{2nk+i}$ the corresponding basis obtained from $\mathfrak E_i$. Then $\mathcal L=(\Lambda_i)_i$ is a complete chain of $\mathcal O_{F_\mathcal P}$-lattices in $V$. For $i\in\ZZ$, the dual lattice $\Lambda_i^\vee:=\{x\in V_{\QQ_p}\mathrel{|} \pairt{x}{\Lambda_i}_{\QQ_p}\subset\ZZ_p\}$ of $\Lambda_i$ is given by $\Lambda_{-i}$. Consequently the chain $\mathcal L$ is self-dual. Let $i\in\ZZ$. We denote by $\rho_i:\Lambda_{i}\to \Lambda_{i+1}$ the inclusion, by $\vartheta_i:\Lambda_{2n+i}\to \Lambda_i$ the isomorphism given by multiplication with $\pi$ and by $\left(\cdot,\cdot\right)_i:\Lambda_i\times \Lambda_{-i}\to \ZZ_p$ the restriction of $\left(\cdot,\cdot\right)_{\QQ_p}$. Then $(\Lambda_{i},\rho_{i},\vartheta_{i},\left(\cdot,\cdot\right)_{i})_i$ is a polarized chain of $\mathcal O_{F_\mc P}$-modules of type $(\mathcal L)$, which, by abuse of notation, we also denote by $\mathcal L$. Denote by $\left<\cdot,\cdot\right>_i:\Lambda_i\times \Lambda_{-i}\to \mathcal O_{F_\mathcal P}$ the restriction of the pairing $\delta^{-1}\left(\cdot,\cdot\right)'_{\QQ_p}$. It is the perfect pairing described by the matrix $\widetilde{J}_{2n}$ with respect to the bases $\mathfrak E_i$ and $\mathfrak E_{-i}$. \subsection{The determinant morphism} Denote by $\Sigma$ the set of all embeddings $F\hookrightarrow \CC$. The canonical isomorphism \begin{equation}\label{Fdecomp} F\otimes\CC = \prod_{\sigma\in\Sigma}\CC \end{equation} induces a decomposition $V\otimes\CC = \prod_{\sigma\in\Sigma} V_\sigma$ into $\CC$-vector spaces $V_\sigma$, and the morphism $J_\CC$ decomposes into the product of $\CC$-linear maps $J_\sigma:V_\sigma\to V_\sigma$. Each $J_\sigma$ induces a decomposition $V_\sigma=V_{\sigma,i}\oplus V_{\sigma,-i}$, where $V_{\sigma,\pm i}$ denotes the $\pm i$-eigenspace of $J_\sigma$. From the explicit description of $J$ in terms of $\mathcal B$ above one sees that both $V_{\sigma,i}$ and $V_{\sigma,-i}$ have dimension $n$ over $\CC$. The $(-i)$-eigenspace $V_{-i}$ of $J_\CC$ is given by $V_{-i}=\prod_{\sigma\in\Sigma} V_{-i,\sigma}$. As $\dim_\CC V_{-i,\sigma}\allowbreak=n$ for all $\sigma$, there is an isomorphism $V_{-i}\simeq (\prod_{\sigma}\CC)^n$ of $\prod_{\sigma}\CC$-modules and hence the $\mathcal O_F\otimes \CC$-module corresponding to $V_{-i}$ under \eqref{Fdecomp} is isomorphic to $\mathcal O_F^n\otimes \CC$. In particular, the morphism $\det_{V_{-i}}:V_{\mathcal O_F\otimes\CC}\to \mathbb{A}^1_\CC$ is defined over $\ZZ$. \subsection{The local model} For the chosen PEL datum, Definition \ref{DefnLocalModelGen} amounts to the following. \begin{defn}\label{DefnLocalModel} The local model $M^{\mathrm{loc}}$ is the functor on the category of $\ZZ_p$-algebras with $M^{\mathrm{loc}}(R)$ the set of tuples $(t_i)_{i\in \ZZ}$ of $\mathcal O_{F}\otimes R$-submodules $t_i\subset \Lambda_{i,R}$ satisfying the following conditions for all $i\in\ZZ$. \renewcommand\theenumi {\alph{enumi}} \begin{enumerate} \item\label{DefnLocalModel-Functoriality} $\rho_{i,R}(t_i)\subset t_{i+1}$. \item\label{DefnLocalModel-Projectivity} The quotient $\Lambda_{i,R}/t_i$ is a finite locally free $R$-module. \item\label{DefnLocalModel-Determinant} We have an equality \begin{equation*} \det_{\Lambda_{i,R}/t_i}=\det_{V_{-i}}\otimes R \end{equation*} of morphisms $V_{\mathcal O_{F}\otimes R}\to \mathbb{A}^1_R$. \item\label{DefnLocalModel-DualityCondition} Under the pairing $\left(\cdot,\cdot\right)_{i,R}:\Lambda_{i,R}\times \Lambda_{-i,R}\to R$, the submodules $t_i$ and $t_{-i}$ pair to zero. \item\label{DefnLocalModel-Periodicity} $\vartheta_i(t_{2n+i})=t_i$. \end{enumerate} \end{defn} \subsection{The special fiber of the local model} For $i\in\ZZ$, denote by $\overline{\Lambda}_i$ the $\mathbb{F}[u]/u^e$-module $(\mathbb{F}[u]/u^e)^{2n}$ and by $\overline{\mathfrak E}_i$ its canonical basis. Denote by $\overline{\left<\cdot,\cdot\right>}_i:\overline{\Lambda}_i\times \overline{\Lambda}_{-i}\to \mathbb{F}[u]/u^e$ the pairing described by the matrix $\widetilde{J}_{2n}$ with respect to $\overline{\mathfrak E}_i$ and $\overline{\mathfrak E}_{-i}$. Denote by $\overline{\vartheta}_i:\overline{\Lambda}_{2n+i}\to \overline{\Lambda}_i$ the identity morphism. For $k\in\ZZ$ and $0\leq i<2n$, let $\overline{\rho}_{2n+i}:\overline{\Lambda}_{2n+i}\to \overline{\Lambda}_{2n+i+1}$ be the morphism described by the matrix $\mathrm{diag}(1^{(i)},u,1^{(2n-i-1)})$ with respect to $\overline{\mathfrak E}_{2nk+i}$ and $\overline{\mathfrak E}_{2nk+i+1}$. \begin{defn}\label{DefnSpecialLocalModel} Let $M^{e,n}$ be the functor on the category of $\mathbb{F}$-algebras with $M^{e,n}(R)$ the set of tuples $(t_i)_{i\in \ZZ}$ of $R[u]/u^e$-submodules $t_i\subset \overline{\Lambda}_{i,R}$ satisfying the following conditions for all $i\in\ZZ$. \renewcommand\theenumi {\alph{enumi}} \begin{enumerate} \item $\overline{\rho}_{i,R}(t_i)\subset t_{i+1}$. \item The quotient $\overline{\Lambda}_{i,R}/t_i$ is finite locally free over $R$. \item\label{DefnSpecialLocalModel-Determinant} For all $p\in R[u]/u^e$, we have \begin{equation*} \chi_R(p|\overline{\Lambda}_{i,R}/t_i)=\bigl(T-p(0)\bigr)^{ne} \end{equation*} in $R[T]$. \item $t_i^{\perp,\overline{\left<\cdot,\cdot\right>}_{i,R}}=t_{-i}$. \item $\overline{\vartheta}_i(t_{2n+i})=t_i$. \end{enumerate} \end{defn} Denote by $\mathfrak S$ the set of all embeddings $k_{\mathcal P}\hookrightarrow \mathbb{F}$. Our choice of uniformizer $\pi$ induces a canonical isomorphism \begin{equation} \label{Eqdecompspecialfiber} \mathcal O_F\otimes\mathbb{F}=\prod_{\sigma\in\mathfrak S} \mathbb{F}[u]/(u^e). \end{equation} Let $i\in\ZZ$. From \eqref{Eqdecompspecialfiber} we obtain an isomorphism \begin{equation}\label{EqLambdaInSpecialFiber} \Lambda_{i,\mathbb{F}}=\prod_{\sigma\in\mathfrak S}\overline{\Lambda}_{i} \end{equation} by identifying the basis $\mathfrak E_{i,\mathbb{F}}$ with the product of the bases $\overline{\mathfrak E}_i$. Under this identification, the morphism $\rho_{i,\mathbb{F}}$ decomposes into the morphisms $\overline{\rho}_i$, the pairing $\left<\cdot,\cdot\right>_{i,\mathbb{F}}$ decomposes into the pairings $\overline{\left<\cdot,\cdot\right>}_i$ and the morphism $\vartheta_{i,\mathbb{F}}$ decomposes into the morphisms $\overline{\vartheta}_i$. Let $R$ be an $\mathbb{F}$-algebra and let $(t_i)_{i\in\ZZ}$ be a tuple of $\mathcal O_{F}\otimes R$-submodules $t_i\subset \Lambda_{i,R}$. Then \eqref{EqLambdaInSpecialFiber} induces decompositions $t_i=\prod_{\sigma\in\mathfrak S} t_{i,\sigma}$ into $R[u]/u^e$-submodules $t_{i,\sigma}\subset\overline{\Lambda}_{i,R}$. \begin{prop}\label{PropDecompositionofLocalModel} The morphism $M^{\mathrm{loc}}_\mathbb{F}\to \prod_{\sigma\in\mathfrak S} M^{e,n}$ given on $R$-valued points by \begin{equation}\label{EqDecompLocModelSym} \begin{aligned} M^{\mathrm{loc}}_\mathbb{F}(R)&\to \prod_{\sigma\in\mathfrak S} M^{e,n}(R),\\ (t_i)&\mapsto \left((t_{i,\sigma})_i\right)_\sigma \end{aligned} \end{equation} is an isomorphism of functors on the category of $\mathbb{F}$-algebras. \end{prop} \begin{proof} The only point requiring an argument is the transition from $\left(\cdot,\cdot\right)_i$ to $\left<\cdot,\cdot\right>_i$. It is warranted by the perfectness of the pairing $\mathcal O_{F_\mc P}\times\mathcal O_{F_\mc P}\to \ZZ_p,\ (x,y)\mapsto \tr_{F_\mc P/\QQ_p}(\delta xy)$. \end{proof} \subsection{The affine Grassmannian and the affine flag variety for \texorpdfstring{$\mathrm{GL}_n$}{GL(n)}}\label{SecLattices} Let $R$ be an $\mathbb{F}$-algebra and let $n\in\NN$. \begin{defn}\label{DefnLattice} A \emph{lattice} in $R(\!(u)\!)^{n}$ is an $R[\![u]\!]$-submodule $L\subset R(\!(u)\!)^{n}$ satisfying the following conditions for some $N\in\NN$. \begin{enumerate} \item\label{DefnLattice-Inclusion} $u^NR[\![u]\!]^{n}\subset L\subset u^{-N}R[\![u]\!]^{n}$. \item\label{DefnLattice-Quotient} $u^{-N}R[\![u]\!]^{n}/L$ is a finite locally free $R$-module. \end{enumerate} \end{defn} The following statement is well-known. See for example \cite[Proposition 4.5.5]{diss} for a proof. \begin{prop}\label{PropLatticesFreeGlobal} Let $L$ be a lattice in $R(\!(u)\!)^n$. Then $L$ is a finite locally free $R[\![u]\!]$-module of rank $n$. \end{prop} \begin{defn} The affine Grassmannian $\mathcal G$ is the functor on the category of $\mathbb{F}$-algebras with $\mathcal G(R)$ the set of lattices in $R(\!(u)\!)^n$. \end{defn} Denote by $\widetilde{\Lambda}_0=R[\![u]\!]^n$ the \emph{standard lattice}. Clearly $\Lf \mathrm{GL}_n(R)$ acts on $\mathcal G(R)$ by multiplication from the left, and the stabilizer of $\widetilde{\Lambda}_0$ for this action is given by $\Lp \mathrm{GL}_n(R)$. Consequently we get an injective map \begin{align*} \phi(R):\Lf \mathrm{GL}_n(R)/\Lp \mathrm{GL}_n(R)&\to \mathcal G(R)\\ g&\mapsto g\widetilde{\Lambda}_0. \end{align*} It is equivariant for the left action by $\Lf \mathrm{GL}_n$. \begin{prop}\label{PropGrassandGLn} The map $\phi$ identifies $\mathcal G$ with both the Zariski and the fpqc sheafification of the presheaf $\Lf \mathrm{GL}_n/\Lp \mathrm{GL}_n$. \end{prop} \begin{proof} By Proposition \ref{PropLatticesFreeGlobal} it is clear that any lattice lies in the image of $\phi$ Zariski locally on $R$. It follows that $\phi$ is the Zariski sheafification of the presheaf $\Lf \mathrm{GL}_n/\Lp \mathrm{GL}_n$. The fact that $\mathcal G$ is already an fpqc sheaf implies formally that $\phi$ is also the fpqc sheafification of the presheaf $\Lf \mathrm{GL}_n/\Lp \mathrm{GL}_n$. \end{proof} \begin{defn} A (complete, periodic) \emph{lattice chain} in $R(\!(u)\!)^n$ is a tuple $(L_i)_{i\in\ZZ}$ of lattices $L_i$ in $R(\!(u)\!)^{n}$ satisfying the following conditions for each $i\in\ZZ$. \begin{enumerate} \item $L_i\subset L_{i+1}$. \item (completeness) $L_{i+1}/L_i$ is a locally free $R$-module of rank $1$. \item (periodicity) $L_{n+i}=u^{-1} L_i$. \end{enumerate} \end{defn} \begin{defn} The \emph{affine flag variety} $\mathcal F$ is the functor on the category of $\mathbb{F}$-algebras with $\mathcal F(R)$ the set of (complete, periodic) lattice chains in $R(\!(u)\!)^{n}$. \end{defn} Denote by $(e_1,\ldots,e_{n})$ the standard basis of $R(\!(u)\!)^{n}$ over $R(\!(u)\!)$. For $0\leq i< n$ we denote by $\widetilde{\Lambda}_i$ the lattice in $R(\!(u)\!)^{n}$ with basis \begin{equation*} \widetilde{\mathfrak E}_i=\left<u^{-1}e_1,\ldots,u^{-1}e_i,e_{i+1},\ldots,e_{n}\right>. \end{equation*} For $k\in\ZZ$ we further define $\widetilde{\Lambda}_{nk+i}=u^{-k}\widetilde{\Lambda}_i$ and we denote by $\widetilde{\mathfrak E}_{nk+i}$ the corresponding basis obtained from $\widetilde{\mathfrak E_i}$. Then $\widetilde{\mathcal L}=(\widetilde{\Lambda}_i)_i$ is a (complete, periodic) lattice chain in $R(\!(u)\!)^n$, called the \emph{standard lattice chain}. In complete analogy with \cite[p. 131]{rz}, we have for an $\mathbb{F}[\![u]\!]$-algebra $R$ the notion of a chain $\mathcal M=(M_i,\varrho_i:M_i\to M_{i+1},\theta_i:M_{n+i}\xrightarrow{\sim} M_i)_{i\in\ZZ}$ of $R$-modules of type $(\widetilde{\mathcal L})$ (cf.\ \cite[Definition 7.5.1]{diss}). The proof of \cite[Proposition A.4]{rz} then carries over without any changes to show the following result. \begin{prop}\label{PropRZbigLin} Let $R$ be an $\mathbb{F}[\![u]\!]$-algebra such that the image of $u$ in $R$ is nilpotent. Then any two chains $\mathcal M,\mathcal N$ of $R$-modules of type $(\widetilde{\mathcal L})$ are isomorphic locally for the Zariski topology on $R$. Furthermore the functor $\Isom(\mathcal M,\mathcal N)$ is representable by a smooth affine scheme over $R$. \end{prop} \begin{prop}\label{PropLiftingIsosLin} Let $R$ be an $\mathbb{F}$-algebra and let $\mathcal M,\mathcal N$ be chains of $R[\![u]\!]$-modules of type $(\widetilde{\mathcal L})$. Then the canonical map $\Isom(\mathcal M,\mathcal N)(R[\![u]\!])\to \Isom(\mathcal M,\mathcal N)(R[\![u]\!]/u^m)$ is surjective for all $m\in\NN_{\geq 1}$. In particular $\mathcal M$ and $\mathcal N$ are isomorphic locally for the Zariski topology on $R$. \end{prop} \begin{proof} Analogous to the proof of Lemma \ref{LemIsomOverK}. \end{proof} \begin{remark}\label{RemarkLatticesAreChainsLin} Let $R$ be an $\mathbb{F}$-algebra and let $(L_i)_i\in \mathcal F(R)$. For $i\in\ZZ$ denote by $\varrho_i:L_i\to L_{i+1}$ the inclusion and by $\theta_i:L_{n+i}\to L_i$ the isomorphism given by multiplication with $u$. Then $(L_i,\varrho_i,\theta_i)$ is a chain of $R[\![u]\!]$-modules of type $(\widetilde{\mathcal L})$. \end{remark} \begin{remark}\label{RemarkActionOfGLOnF} The group $\Lf \mathrm{GL}_n(R)$ acts on $\mathcal F(R)$ via $g\cdot (L_i)_i=(gL_i)_i$. Denote by $I(R)$ the stabilizer of $\widetilde{\mathcal L}$ for this action. One checks that $I(R)\subset \mathrm{GL}_n(R[\![u]\!])$ is equal to the preimage of $B(R)$ under the reduction map $\mathrm{GL}_n(R[\![u]\!])\to \mathrm{GL}_n(R),\ u\mapsto 0$. Here $B(R)\subset \mathrm{GL}_n(R)$ denotes the subgroup of upper triangular matrices. \end{remark} We obtain for each $\mathbb{F}$-algebra $R$ an injective map \begin{align*} \Lf {\mathrm{GL}_n}(R)/I(R)&\xrightarrow{\phi(R)} \mathcal F(R),\\ g&\xmapsto{\hphantom{\phi(R)}} g\cdot \widetilde{\mathcal L}. \end{align*} \begin{prop}\label{PropFlagandGL} The morphism $\phi$ identifies $\mathcal F$ with both the Zariski and the fpqc sheafification of the presheaf $\Lf \mathrm{GL}_n/I$. \end{prop} \begin{proof} Let $R$ be an $\mathbb{F}$-algebra and let $\mathcal M\in\mathcal F(R)$. We consider $\mathcal M$ as a chain of $R[\![u]\!]$-modules of type $(\widetilde{\mathcal L})$ as in Remark \ref{RemarkLatticesAreChainsLin}. By Proposition \ref{PropLiftingIsosLin}, the chains $\widetilde{\mathcal L}$ and $\mathcal M$ are isomorphic locally for the Zariski topology on $R$. Such an isomorphism $\widetilde{\mathcal L} \to \mathcal M$ is given by multiplication with a single $g\in \mathrm{GL}_n(R[\![u]\!])$. Consequently $\mathcal M$ lies in the image of $\phi$ Zariski locally on $R$. The fact that $\mathcal F$ is already an fpqc sheaf implies formally that $\phi$ is also the fpqc sheafification of the presheaf $\Lf \mathrm{GL}_n/I$. \end{proof} \subsection{The affine flag variety}\label{SecAffFlagSym} This section deals with the affine flag variety for the symplectic group. Our discussion loosely follows the one in \cite[\textsection 10-11]{pr2}. Note though that in loc.\ cit.\ there is a minor problem with the definition of the notion of self-duality for lattice chains, see Remark \ref{RemWrongDualDefn} below. We have learned the correct formulation of this definition from \cite[\textsection 4.2]{smithling_unitary_odd}, which deals with the case of a ramified unitary group. Let $R$ be an $\mathbb{F}$-algebra. Let $\widetilde{\left<\cdot,\cdot\right>}$ be the symplectic form on $R(\!(u)\!)^{2n}$ described by the matrix $\widetilde{J}_{2n}$ with respect to the standard basis of $R(\!(u)\!)^{2n}$ over $R(\!(u)\!)$. We denote by $\mathrm{Sp}=\mathrm{Sp}_{2n}$ the symplectic group and by $\mathrm{GSp}=\mathrm{GSp}_{2n}$ the group of symplectic similitudes with respect to $\widetilde{\left<\cdot,\cdot\right>}$. For a lattice $\Lambda$ in $R(\!(u)\!)^{2n}$ we define $\Lambda^\vee:=\{x\in R(\!(u)\!)^{2n}\mathrel{|} \widetilde{\pair{x}{\Lambda}}\subset R[\![u]\!]\}$. Recall from Section \ref{SecLattices} the standard lattice chain $\widetilde{\mathcal L}=(\widetilde{\Lambda}_i)_i$ in $R(\!(u)\!)^{2n}$. Note that $(\widetilde{\Lambda}_i)^\vee=\widetilde{\Lambda}_{-i}$ for all $i\in\ZZ$. We denote by $\widetilde{\left<\cdot,\cdot\right>}_i:\widetilde{\Lambda}_i\times \widetilde{\Lambda}_{-i}\to R[\![u]\!]$ the restriction of $\widetilde{\left<\cdot,\cdot\right>}$. In complete analogy with \cite[Definition 3.14]{rz}, we have for an $\mathbb{F}[\![u]\!]$-algebra $R$ the notion of a polarized chain $\mathcal M=(M_i,\varrho_i:M_i\to M_{i+1},\theta_i:M_{2n+i}\xrightarrow{\sim} M_i, \mathcal E_i:M_i\times M_{-i}\to R)_{i\in\ZZ}$ of $R$-modules of type $(\widetilde{\mathcal L})$ (cf.\ \cite[Definition 5.5.1]{diss}). The proof of \cite[Proposition A.21]{rz} then carries over without any changes the show the following result. \begin{prop}\label{PropRZbig} Let $R$ be an $\mathbb{F}[\![u]\!]$-algebra such that the image of $u$ in $R$ is nilpotent. Then any two polarized chains $\mathcal M,\mathcal N$ of $R$-modules of type $(\widetilde{\mathcal L})$ are isomorphic locally for the Zariski topology on $R$. Furthermore the functor $\Isom(\mathcal M,\mathcal N)$ is representable by a smooth affine scheme over $R$. \end{prop} \begin{prop}\label{PropLiftingIsosSym} Let $R$ be an $\mathbb{F}$-algebra and let $\mathcal M,\mathcal N$ be polarized chains of $R[\![u]\!]$-modules of type $(\widetilde{\mathcal L})$. Then the canonical map $\Isom(\mathcal M,\mathcal N)(R[\![u]\!])\to \Isom(\mathcal M,\mathcal N)(R[\![u]\!]/u^m)$ is surjective for all $m\in\NN_{\geq 1}$. In particular $\mathcal M$ and $\mathcal N$ are isomorphic locally for the Zariski topology on $R$. \end{prop} \begin{proof} Analogous to the proof of Lemma \ref{LemIsomOverK}. \end{proof} The following definition is a straightforward variant of \cite[\textsection 4.2]{smithling_unitary_odd}. \begin{defn} Let $R$ be an $\mathbb{F}$-algebra and let $(L_i)_i$ be a lattice chain in $R(\!(u)\!)^{2n}$. \begin{enumerate} \item Let $r\in\ZZ$. The chain $(L_i)_i$ is called \emph{$r$-self-dual} if \begin{equation*} \forall i\in\ZZ:\ L_i^\vee=u^r L_{-i}. \end{equation*} Denote by $\mathcal F^{(r)}_\mathrm{Sp}$ the functor on the category of $\mathbb{F}$-algebras with $\mathcal F_{\mathrm{Sp}}^{(r)}(R)$ the set of $r$-self-dual lattice chains in $R(\!(u)\!)^{2n}$. \item The chain $(L_i)_i$ is called \emph{self-dual} if Zariski locally on $R$ there is an $a\in R(\!(u)\!)^\times$ such that \begin{equation}\label{EquSelfdual} \forall i\in\ZZ:\ L_i^\vee=a L_{-i}. \end{equation} We denote by $\mathcal F_{\mathrm{GSp}}$ the functor on the category of $\mathbb{F}$-algebras with $\mathcal F_{\mathrm{GSp}}(R)$ the set of self-dual lattice chains in $R(\!(u)\!)^{2n}$. \end{enumerate} \end{defn} Note that $\widetilde{\mathcal L}\in \mathcal F_\mathrm{Sp}^{(0)}(R)$. \begin{lem}\label{LemUnitsinLaurent} Let $R$ be a ring and let $a\in R(\!(u)\!)^\times$. Then Zariski locally on $R$, there are integers $n\leq n_0$, nilpotent elements $a_{n},a_{n+1},\dots,a_{n_0-1}\in R$, a unit $a_{n_0}\in R^\times$ and elements $a_{n_0+1},a_{n_0+2},\ldots\in R$ such that $a=\sum_{i=n}^{\infty}a_i u^i$. If $\Spec R$ is connected, such integers and elements exist globally on $R$. \end{lem} \begin{remark}\label{RemarkLocalModel} Let $R$ be a reduced $\mathbb{F}$-algebra such that $\Spec R$ connected. Then \begin{equation*} \mathcal F_{\mathrm{GSp}}(R)=\bigcup_{r\in\ZZ} \mathcal F_{\mathrm{Sp}}^{(r)}(R). \end{equation*} \end{remark} \begin{proof} This follows immediately from Lemma \ref{LemUnitsinLaurent}. \end{proof} \begin{remark}\label{RemarkLatticesAreChainsSym} Let $R$ be an $\mathbb{F}$-algebra and let $(L_i)_i\in \mathcal F_{\mathrm{Sp}}^{(0)}(R)$. For $i\in\ZZ$ denote by $\varrho_i:L_i\to L_{i+1}$ the inclusion, by $\theta_i:L_{2n+i}\to L_i$ the isomorphism given by multiplication with $u$ and by $\mathcal E_i:L_i\times L_{-i}\to R[\![u]\!]$ the restriction of $\widetilde{\left<\cdot,\cdot\right>}$. Then $(L_i,\varrho_i,\theta_i,\mathcal E_i)$ is a polarized chain of $R[\![u]\!]$-modules of type $(\widetilde{\mathcal L})$. \end{remark} Recall from Remark \ref{RemarkActionOfGLOnF} the subfunctor $I\subset \Lf \mathrm{GL}_{2n}$. We define a subfunctor $I_{\mathrm{GSp}}=I_{\mathrm{GSp}_{2n}}$ of $\Lf \mathrm{GSp}=\Lf \mathrm{GSp}_{2n}$ by $I_\mathrm{GSp}=\Lf \mathrm{GSp}_{2n}\cap I$. We consider all of these functors as functors on the category of $\mathbb{F}$-algebras. The proof of the following result is similar to and therefore based on the proof of \cite[Theorem 4.1]{pr}. \begin{prop}\label{PropFuncDescofFlag} The natural action of $\Lf \mathrm{GL}_{2n}$ on $\mathcal F$ (cf.\ Remark \ref{RemarkActionOfGLOnF}) restricts to an action of $\Lf \mathrm{GSp}$ on $\mathcal F_\mathrm{GSp}$. Consequently we obtain an injective map \begin{align*} \Lf {\mathrm{GSp}}(R)/I_{\mathrm{GSp}}(R)&\xrightarrow{\phi(R)} \mathcal F_{\mathrm{GSp}}(R),\\ g&\xmapsto{\hphantom{\phi(R)}} g\cdot \widetilde{\mathcal L} \end{align*} for each $\mathbb{F}$-algebra $R$. The morphism $\phi$ identifies $\mathcal F_{\mathrm{GSp}}$ with both the Zariski and the fpqc sheafification of the presheaf $\Lf {\mathrm{GSp}}/I_{\mathrm{GSp}}$. \end{prop} \begin{proof} Let $R$ be an $\mathbb{F}$-algebra and let $\mathcal M=(L_i)_i\in\mathcal F_{\mathrm{GSp}}(R)$. Working Zariski locally on $R$ we may assume that there is an $a\in R(\!(u)\!)^\times$ such that \eqref{EquSelfdual} holds. Choose any $h\in {\mathrm{GSp}}(R(\!(u)\!))$ with factor of similitude $a$, e.g.\ $h=\diag(a^{(n)},1^{(n)})$. An easy computation shows that $h\mathcal M\in \mathcal F_{\mathrm{Sp}}^{(0)}(R)$. We see as in the proof of Proposition \ref{PropFlagandGL} that Zariski locally on $R$, there is a $g\in \mathrm{Sp}(R(\!(u)\!))$ with $h\mathcal M=g\widetilde{\mathcal L}$. Consequently $\mathcal M=h^{-1}g\widetilde{\mathcal L}$ lies in the image of $\phi$ Zariski locally on $R$. As $\mathcal F_{\mathrm{GSp}}$ is clearly a Zariski sheaf, it follows that $\mathcal F_{\mathrm{GSp}}$ is indeed the Zariski sheafification of the presheaf $\Lf {\mathrm{GSp}}/I_{\mathrm{GSp}}$. To see that $\mathcal F_{\mathrm{GSp}}$ is also the fpqc sheafification of $\Lf {\mathrm{GSp}}/I_{\mathrm{GSp}}$, it suffices to show that $\mathcal F_{\mathrm{GSp}}$ is an fpqc sheaf. Let $(L_i)_i$ be a lattice chain in $R(\!(u)\!)^{2n}$. Assume that fpqc locally on $R$ there is an $a\in R(\!(u)\!)^\times$ such that \eqref{EquSelfdual} holds. The scalar $a$ gives rise to a well-defined element of $(\Lf \GG_m/\Lp \GG_m)_{\mathrm{fpqc}}(R)$, where $(\Lf \GG_m/\Lp \GG_m)_{\mathrm{fpqc}}$ denotes the fpqc sheafification of the presheaf $\Lf \GG_m/\Lp \GG_m$. By Proposition \ref{PropGrassandGLn} any element of $(\Lf \GG_m/\Lp \GG_m)_{\mathrm{fpqc}}(R)$ can be represented in $\Lf \GG_m(R)$ Zariski locally on $R$, so that the scalar $a$ exists in fact Zariski locally on $R$. \end{proof} \begin{remark}\label{RemWrongDualDefn} Let us note that there seems to exist a misconception surrounding the notion of self-duality for lattice chains. In the literature one finds the following definition: Let $R$ be an $\mathbb{F}$-algebra. A lattice chain $(L_i)_i\in \mathcal F(R)$ is called \emph{(naively) self-dual} if for each $i\in\ZZ$ there is a $j\in\ZZ$ such that $L_{i}^\vee= L_j$. It is then claimed that the fpqc local $\Lf {\mathrm{GSp}}$-orbit of $\widetilde{\mathcal L}$ (in the sense of Proposition \ref{PropFuncDescofFlag}) is precisely the set of (naively) self-dual lattice chains. This is wrong in both directions, as shown by the following easy examples. \begin{itemize}[leftmargin=*] \item Let $n=1$ and $a\in R(\!(u)\!)^\times$. The chain $(L_i)_i=a\widetilde{\mathcal L}$ satisfies $L_i^\vee=a^{-2}L_{-i},\ i\in\ZZ$. Assume there is a $j\in\ZZ$ with $L_0^\vee=L_j$. Then $a^{-2}L_{0}=L_j$ and hence $a^{-2}\widetilde{\Lambda}_{0}=\widetilde{\Lambda}_j$. Projecting this equality inside $R(\!(u)\!)^2$ to its first components yields the existence of a $k\in\ZZ$ with $a^{-2}R[\![u]\!]=u^{k}R[\![u]\!]$, so that $u^ka^2\in R[\![u]\!]^\times$. If for example $R=\mathbb{F}[x]/x^2$ and $a=1+xu^{-1}$, such a $k$ does not exist. \item Conversely one easily sees that for $n\geq 2$, the (naively) self-dual chain $(\widetilde{\Lambda}_{i+1})_{i\in\ZZ}$ does not lie in the $\Lf \mathrm{GSp}(R)$-orbit of $\widetilde{\mathcal L}$ (unless $R=\{0\}$). \end{itemize} \end{remark} \subsection{Embedding the local model into the affine flag variety}\label{SecEmbLocFlagSym} Let $R$ be an $\mathbb{F}$-algebra. We consider an $R[u]/u^e$-module as an $R[\![u]\!]$-module via the canonical projection $R[\![u]\!]\to R[u]/u^e$. For $i\in\ZZ$ denote by $\alpha_i:\widetilde{\Lambda}_i\to \overline{\Lambda}_{i,R}$ the morphism described by the identity matrix with respect to $\widetilde{\mathfrak E}_i$ and $\overline{\mathfrak E}_i$. It induces an isomorphism $\widetilde{\Lambda}_i/u^e\widetilde{\Lambda}_i\xrightarrow{\sim}\overline{\Lambda}_{i,R}$. Clearly the following diagrams commute. \begin{equation*} \xymatrix{ \widetilde{\Lambda}_i\ar[d]_{\alpha_i}\ar@{}[r]|{\subset}&\widetilde{\Lambda}_{i+1}\ar[d]^{\alpha_{i+1}}\\ \overline{\Lambda}_{i,R}\ar[r]^{\overline{\rho}_{i,R}}&\overline{\Lambda}_{i+1,R}, }\;\; \xymatrix{ \widetilde{\Lambda}_i\times \widetilde{\Lambda}_{-i}\ar[r]^-{\widetilde{\left<\cdot,\cdot\right>}_i}\ar[d]_{\alpha_i\times \alpha_{-i}}& R[\![u]\!]\ar[d]\\ \overline{\Lambda}_{i,R}\times \overline{\Lambda}_{-i,R}\ar[r]^-{\overline{\left<\cdot,\cdot\right>}_{i,R}}& R[u]/u^e, } \;\; \xymatrix{ \widetilde{\Lambda}_i\ar[d]_{\alpha_i}&\widetilde{\Lambda}_{2n+i}\ar[d]^{\alpha_{2n+i}}\ar[l]_{u\cdot }\\ \overline{\Lambda}_{i,R}&\overline{\Lambda}_{2n+i,R}\ar[l]_{\overline{\vartheta}_{i,R}}. } \end{equation*} The following proposition allows us to consider $M^{e,n}$ as a subfunctor of $\mathcal F_{\mathrm{Sp}}^{(-e)}$. \begin{prop}[{\cite[\textsection 11]{pr2}}]\label{PropLocalModelintoFlag} There is an embedding $\alpha:M^{e,n}\hookrightarrow \mathcal F_{\mathrm{Sp}}^{(-e)}$ given on $R$-valued points by \begin{align*} M^{e,n}(R)&\to \mathcal F_{\mathrm{Sp}}^{(-e)}(R),\\ (t_i)_i&\mapsto (\alpha_{i}^{-1}(t_i))_i. \end{align*} It induces a bijection from $M^{e,n}(R)$ onto the set of those $(L_i)_i\in \mathcal F_{\mathrm{Sp}}^{(-e)}(R)$ satisfying the following conditions for all $i\in\ZZ$. \begin{enumerate} \item $u^e\widetilde{\Lambda}_i\subset L_i\subset\widetilde{\Lambda}_i$. \item For all $p\in R[u]/u^e$, we have \begin{equation*} \chi_R(p|\widetilde{\Lambda}_i/L_i)=(T-p(0))^{ne} \end{equation*} in $R[T]$. Here $\widetilde{\Lambda}_i/L_i$ is considered as an $R[u]/u^e$-module using (1). \end{enumerate} \end{prop} \begin{proof} Let $(t_i)_i\in M^{e,n}(R)$ and set $(L_i)_i= (\alpha_i^{-1}(t_i))_i$. It is clear that this defines a periodic lattice chain in $R(\!(u)\!)^{2n}$. Let $i\in\ZZ$. We have \begin{equation}\label{EqRanks} \rk_R(\widetilde{\Lambda}_{i+1}/\widetilde{\Lambda}_i)+\rk_R(\widetilde{\Lambda}_i/L_i)=\rk_R(\widetilde{\Lambda}_{i+1}/L_{i+1})+\rk_R(L_{i+1}/L_i), \end{equation} as both sides are equal to $\rk_R(\widetilde{\Lambda}_{i+1}/L_i)$. We conclude from condition \ref{DefnSpecialLocalModel}(\ref{DefnSpecialLocalModel-Determinant}) that $\rk_R(\widetilde{\Lambda}_i/L_i)=ne=\rk_R(\widetilde{\Lambda}_{i+1}/L_{i+1})$. Thus \eqref{EqRanks} amounts to the equation $\rk_R(\widetilde{\Lambda}_{i+1}/\widetilde{\Lambda}_i)=\rk_R(L_{i+1}/L_i)$, so that the chain $(L_i)_i$ is complete. From $\overline{\pair{t_i}{t_{-i}}}_{i,R}=0$ we deduce that $\widetilde{\pair{\mathcal L_i}{L_{-i}}}\subset u^eR[\![u]\!]$ and hence that $u^{-e}L_{-i}\subset L_i^\vee$. From $u^e\widetilde{\Lambda}_i\subset L_i$ on the other hand we deduce $u^e L_i^\vee\subset \widetilde{\Lambda}_{-i}$. By definition, we know that $\widetilde{\pair{L_i}{u^e L_i^\vee}}\subset u^eR[\![u]\!]$, which implies $\overline{\pair{t_i}{\alpha_{-i}(u^e L_i^\vee)}}_{i}=0$. Consequently $\alpha_{-i}(u^e L_i^\vee)\subset t_{-i}$, which shows that $u^e L_i^\vee \subset L_{-i}$. Hence also $L_i^\vee \subset u^{-e} L_{-i}$. This proves the existence of the map $\alpha$. Its injectivity as well as the characterization of its image are immediate. \end{proof} Note that $\overline{\mathcal L}=(\overline{\Lambda}_{i},\overline{\rho}_{i},\overline{\vartheta}_{i},\overline{\left<\cdot,\cdot\right>}_{i})_i$ is a polarized chain of $\mathbb{F}[u]/u^e$-modules of type $(\widetilde{\mathcal L})$. In fact $\overline{\mathcal L}=\widetilde{\mathcal L}\otimes_{\mathbb{F}[\![u]\!]}\mathbb{F}[u]/u^e$. Let $R$ be an $\mathbb{F}$-algebra. There is an obvious action of $\Aut(\overline{\mathcal L})(R[u]/u^e)$ on $M^{e,n}(R)$, given by $(\varphi_i)\cdot (t_{i})=(\varphi_i(t_i))$. The canonical morphism $R[\![u]\!]\to R[u]/u^e$ induces a morphism $\Aut(\widetilde{\mathcal L})(R[\![u]\!])\to \Aut(\overline{\mathcal L})(R[u]/u^e)$ and we thereby extend this $\Aut(\overline{\mathcal L})(R[u]/u^e)$-action on $M^{e,n}(R)$ to an $\Aut(\widetilde{\mathcal L})(R[\![u]\!])$-action. \begin{lem}\label{LemDifferentAutsSameOrbitsSym} Let $R$ be an $\mathbb{F}$-algebra and let $t\in M^{e,n}(R)$. We have \allowbreak$\Aut(\widetilde{\mathcal L})(R[\![u]\!])\cdot t=\Aut(\overline{\mathcal L})(R[u]/u^e)\cdot t$. \end{lem} \begin{proof} The map $\Aut(\widetilde{\mathcal L})(R[\![u]\!])\to \Aut(\overline{\mathcal L})(R[u]/u^e)$ is surjective by Proposition \ref{PropLiftingIsosSym}. \end{proof} Define a subfunctor $I_\mathrm{Sp}=I_{\mathrm{Sp}_{2n}}$ of $\Lf \mathrm{Sp}_{2n}$ by $I_\mathrm{Sp}=\Lf \mathrm{Sp}_{2n}\cap I_{\mathrm{GSp}}$. \begin{lem}\label{LemIvsI0Sym} We have $I_\mathrm{GSp}(\mathbb{F})=\mathbb{F}[\![u]\!]^\times I_\mathrm{Sp}(\mathbb{F})$. \end{lem} \begin{proof} Let $g\in I_\mathrm{GSp}(\mathbb{F})$. Clearly $c(g)\in \mathbb{F}[\![u]\!]^\times$. As $\ch \mathbb{F}\neq 2$, there is an $x\in \mathbb{F}[\![u]\!]^\times$ with $x^2=c(g)$. Then $x^{-1}g\in I_\mathrm{Sp}(\mathbb{F})$. \end{proof} \begin{lem}\label{LemIvsIprimeSym} Let $g\in I_\mathrm{Sp}(\mathbb{F})$. Then $g$ restricts to an automorphism $g_i:\widetilde{\Lambda}_i\to\widetilde{\Lambda}_i$ for each $i\in\ZZ$. The assignment $g\mapsto (g_i)_i$ defines an isomorphism $I_\mathrm{Sp}(\mathbb{F})\xrightarrow{\sim} \Aut(\widetilde{\mathcal L})(\mathbb{F}[\![u]\!])$. \end{lem} \begin{proof} Analogous to the proof of Lemma \ref{LemIvsIprime}. \end{proof} \begin{prop}\label{PropIndicesSym2} Let $t\in M^{e,n}(\mathbb{F})$. Then $\alpha$ induces a bijection \begin{equation*}\label{EqOrbits1Sym} \Aut(\overline{\mathcal L})(\mathbb{F}[u]/u^e)\cdot t\xrightarrow{\sim} I_{\mathrm{GSp}}(\mathbb{F})\cdot \alpha(t). \end{equation*} Consequently we obtain an embedding \begin{equation*} \Aut(\overline{\mathcal L})(\mathbb{F}[u]/u^e)\backslash M^{e,n}(\mathbb{F})\hookrightarrow I_{\mathrm{GSp}}(\mathbb{F})\backslash \mathcal F_\mathrm{GSp}(\mathbb{F}). \end{equation*} \end{prop} \begin{proof} The composition $M^{e,n}(\mathbb{F})\xrightarrow{\alpha}\mathcal F^{(-e)}_\mathrm{Sp}(\mathbb{F})\subset \mathcal F_\mathrm{GSp}(\mathbb{F})$ is equivariant for the $\Aut(\widetilde{\mathcal L})(\mathbb{F}[\![u]\!])$-action on $M^{e,n}(\mathbb{F})$, the $I_\mathrm{Sp}(\mathbb{F})$-action on $\mathcal F_\mathrm{GSp}(\mathbb{F})$ and the isomorphism $I_\mathrm{Sp}(\mathbb{F})\xrightarrow{\sim} \Aut(\widetilde{\mathcal L})(\mathbb{F}[\![u]\!])$ of Lemma \ref{LemIvsIprimeSym}. It therefore induces a bijection $\Aut(\widetilde{\mathcal L})(\mathbb{F}[\![u]\!])\cdot t\xrightarrow{\sim} I_{\mathrm{Sp}}(\mathbb{F})\cdot \alpha(t)$. We conclude by applying Lemmata \ref{LemDifferentAutsSameOrbitsSym} and \ref{LemIvsI0Sym}. \end{proof} Consider $\alpha':M^{e,n}(\mathbb{F})\hookrightarrow \mathcal F_\mathrm{GSp}(\mathbb{F})\xrightarrow{\phi(\mathbb{F})^{-1}}\Lf {\mathrm{GSp}}(\mathbb{F})/I_{\mathrm{GSp}}(\mathbb{F})$. \begin{prop}\label{PropIndicesSym} Let $t\in M^{e,n}(\mathbb{F})$. Then $\alpha'$ induces a bijection \begin{equation*} \Aut(\overline{\mathcal L})(\mathbb{F}[u]/u^e)\cdot t\xrightarrow{\sim} I_{\mathrm{GSp}}(\mathbb{F})\cdot \alpha'(t). \end{equation*} Consequently we obtain an embedding \begin{equation*} \Aut(\overline{\mathcal L})(\mathbb{F}[u]/u^e)\backslash M^{e,n}(\mathbb{F})\hookrightarrow I_{\mathrm{GSp}}(\mathbb{F})\backslash \mathrm{GSp}(\mathbb{F}(\!(u)\!))/I_{\mathrm{GSp}}(\mathbb{F}). \end{equation*} \end{prop} \begin{proof} Clear from Proposition \ref{PropIndicesSym2}, as the isomorphism $\phi(\mathbb{F})$ is in particular $I_{\mathrm{GSp}}(\mathbb{F})$-equivariant. \end{proof} Let $R$ be an $\mathbb{F}$-algebra and $(\varphi_i)_i\in \Aut(\mathcal L)(R)$. The decomposition \eqref{EqLambdaInSpecialFiber} induces for each $i$ a decomposition of $\varphi_i:\Lambda_{i,R}\xrightarrow{\sim}\Lambda_{i,R}$ into the product of $R[u]/u^e$-linear automorphisms $\varphi_{i,\sigma}:\overline{\Lambda}_{i,R}\xrightarrow{\sim}\overline{\Lambda}_{i,R}$. The following statement is then clear (cf.\ the proof of Proposition \ref{PropDecompositionofLocalModel}). \begin{prop}\label{PropChainMorphismDecomposes} Let $R$ be an $\mathbb{F}$-algebra. The following map is an isomorphism, functorial in $R$. \begin{align*} \Aut(\mathcal L)(R)&\to \prod_{\sigma\in\mathfrak S} \Aut(\overline{\mathcal L})(R[u]/u^e),\\ (\varphi_i)_i&\mapsto ((\varphi_{i,\sigma})_i)_\sigma. \end{align*} \end{prop} Consider the composition \begin{equation*} \tilde{\alpha}:M^{\mathrm{loc}}(\mathbb{F})\xrightarrow{\eqref{EqDecompLocModelSym}} \prod_{\sigma \in\mathfrak S} M^{e,n}(\mathbb{F})\xrightarrow{\prod_\sigma \alpha'} \prod_{\sigma\in\mathfrak S} \Lf {\mathrm{GSp}}(\mathbb{F})/I_{\mathrm{GSp}}(\mathbb{F}). \end{equation*} For $\sigma\in \mathfrak S$ denote by $\tilde{\alpha}_\sigma:M^{\mathrm{loc}}(\mathbb{F})\to \Lf {\mathrm{GSp}}(\mathbb{F})/I_{\mathrm{GSp}}(\mathbb{F})$ the corresponding component of $\tilde{\alpha}$. \begin{thm}\label{ThmIndicesSym} Let $t\in M^{\mathrm{loc}}(\mathbb{F})$. Then $\tilde{\alpha}$ induces a bijection \begin{equation*} \Aut(\mathcal L)(\mathbb{F})\cdot t\xrightarrow{\sim} \prod_{\sigma\in\mathfrak S} I_{\mathrm{GSp}}(\mathbb{F})\cdot \tilde{\alpha}_\sigma(t). \end{equation*} Consequently we obtain an embedding \begin{equation*} \Aut(\mathcal L)(\mathbb{F})\backslash M^{\mathrm{loc}}(\mathbb{F})\hookrightarrow \prod_{\sigma\in\mathfrak S} I_{\mathrm{GSp}}(\mathbb{F})\backslash \mathcal \mathrm{GSp}(F(\!(u)\!))/I_{\mathrm{GSp}}(\mathbb{F}). \end{equation*} \end{thm} \begin{proof} The isomorphism $M^{\mathrm{loc}}(\mathbb{F})\xrightarrow{\eqref{EqDecompLocModelSym}} \prod_{\sigma\in\mathfrak S} M^{e,n}(\mathbb{F})$ is equivariant for the $\Aut(\mathcal L)(\mathbb{F})$ action on $M^{\mathrm{loc}}(\mathbb{F})$, the $\prod_{\sigma\in\mathfrak S} \Aut(\overline{\mathcal L})(\mathbb{F}[u]/u^e)$ action on $\prod_{\sigma\in\mathfrak S} M^{e,n}(\mathbb{F})$ and the isomorphism of Lemma \ref{PropChainMorphismDecomposes}. The statement thus follows from Proposition \ref{PropIndicesSym}. \end{proof} \subsection{The extended affine Weyl group}\label{SecExtendedAffineWeyl} Let $T$ be the maximal torus of diagonal matrices in $\mathrm{GSp}_{2n}$ and let $N$ be its normalizer. We denote by $\widetilde{W}=N(\mathbb{F}(\!(u)\!))/T(\mathbb{F}[\![u]\!])$ the extended affine Weyl group of $\mathrm{GSp}$ with respect to $T$. Setting \begin{equation*} W=\{w\in S_{2n}\mathrel{|} \forall i\in\{1,\dots,2n\}:\ w(i)+w(2n+1-i)=2n+1\} \end{equation*} and \begin{equation*} X=\{(a_1,\dots,a_{2n})\in \ZZ^{2n}\mathrel{|} a_1+a_{2n}=a_2+a_{2n-1}=\dots=a_n+a_{n+1}\}, \end{equation*} the group homomorphism $\upsilon:W\ltimes X\to N(\mathbb{F}(\!(u)\!)),\ (w,\lambda)\mapsto A_wu^\lambda$ induces an isomorphism $W\ltimes X\xrightarrow{\sim} \widetilde{W}$. We use it to identify $\widetilde{W}$ with $W\ltimes X$ and consider $\widetilde{W}$ as a subgroup of $\mathrm{GSp}(\mathbb{F}(\!(u)\!)\!)$ via $\upsilon$. To avoid any confusion of the product inside $\widetilde{W}$ and the canonical action of $S_{2n}$ on $\ZZ^{2n}$, we will always denote the element of $\widetilde{W}$ corresponding to $\lambda\in X$ by $u^\lambda$. Recall from \cite[\textsection 2.5-2.6]{gy1} the notion of an extended alcove $(x_i)_{i=0}^{2n-1}$ for $\mathrm{GSp}_{2n}$. Also recall the standard alcove $(\omega_i)_{i=0}^{2n-1}$. As in loc.\ cit.\ we identify $\widetilde{W}$ with the set of extended alcoves by using the standard alcove as a base point. Write $\mathbf{e}=(e^{(2n)})$. \begin{defn}[{Cf.\ \cite{kottwitz_rapoport}, \cite[Definition 2.4]{gy1}}]\label{DefnPermSym} An extended alcove $(x_i)_{i=0}^{2n-1}$ is called \emph{permissible} if it satisfies the following conditions for all $i\in\{0,\dots,2n-1\}$. \begin{enumerate} \item $\omega_i\leq x_i \leq \omega_i+\mathbf{e}$, where $\leq$ is to be understood componentwise. \item $\sum_{j=1}^{2n} x_i(j)=ne-i$. \end{enumerate} Denote by $\mathrm{Perm}$ the set of all permissible extended alcoves. \end{defn} \begin{prop} The inclusion $N(\mathbb{F}(\!(u)\!))\subset \mathrm{GSp}(\mathbb{F}(\!(u)\!))$ induces a bijection $\widetilde{W}\xrightarrow{\sim} I_{\mathrm{GSp}}(\mathbb{F})\backslash \mathcal \mathrm{GSp}(\mathbb{F}(\!(u)\!))/I_{\mathrm{GSp}}(\mathbb{F})$. In other words, \begin{equation*} \mathrm{GSp}(\mathbb{F}(\!(u)\!))=\coprod_{x\in\widetilde{W}} I_{\mathrm{GSp}}(\mathbb{F})xI_{\mathrm{GSp}}(\mathbb{F}). \end{equation*} Under this bijection, the subset \begin{equation*} \Aut(\overline{\mathcal L})(\mathbb{F}[u]/u^e)\backslash M^{e,n}(\mathbb{F})\subset I_{\mathrm{GSp}}(\mathbb{F})\backslash \mathrm{GSp}(\mathbb{F}(\!(u)\!))/I_{\mathrm{GSp}}(\mathbb{F}) \end{equation*} of Proposition \ref{PropIndicesSym} corresponds to the subset $\mathrm{Perm}\subset \widetilde{W}$. \end{prop} \begin{proof} The first statement is the well-known Iwahori decomposition. The second statement follows easily from the explicit description of the image of $\alpha$ in Proposition \ref{PropLocalModelintoFlag}. \end{proof} \begin{cor}\label{CorIndexSetSym} Under the identifications of Theorem \ref{ThmIndicesSym}, the set \allowbreak$\prod_{\sigma\in\mathfrak S} \mathrm{Perm}$ constitutes a set of representatives of $\Aut(\mathcal L)(\mathbb{F})\backslash M^{\mathrm{loc}}(\mathbb{F})$. \end{cor} \begin{remark}\label{RemVvsFSym} In the normalization of Corollary \ref{CorIndexSetSym} we have indexed the KR stratification by the relative position of $\mathcal L\otimes \mathbb{F}$ to its image under \emph{Verschiebung}, compare Remark \ref{RemarkVvsF}. The normalization of Corollary \ref{CorIndexSetSym} therefore differs from the one of Definition \ref{DefnIndexKR} by the automorphism $\prod_{\sigma\in\mathfrak S} \mathrm{Perm}\to \prod_{\sigma\in\mathfrak S} \mathrm{Perm},\ (x_\sigma)\mapsto (u^ex_\sigma^{-1})$. \end{remark} \subsection{The \texorpdfstring{$p$}{p}-rank on a KR stratum} Recall from Section \ref{SecLocalModelGen} the scheme $\mathcal A/\ZZ_p$ associated with our choice of PEL datum, and the KR stratification \begin{equation*} \mathcal A(\mathbb{F})=\coprod_{x\in \Aut(\mathcal L)(\mathbb{F})\backslash M^{\mathrm{loc}}(\mathbb{F})}\mathcal A_x. \end{equation*} We have identified the occurring index set with $\prod_{\sigma\in\mathfrak S}\mathrm{Perm}$ in Corollary \ref{CorIndexSetSym}. We can then state the following result. \begin{thm}\label{ThmPrank} Let $x=(x_\sigma)_\sigma\in \prod_{\sigma\in\mathfrak S} \mathrm{Perm}$. Write $x_\sigma=w_\sigma u^{\lambda_\sigma}$ with $w_\sigma\in W,\ \lambda_\sigma\in X$. Then the $p$-rank on $\mathcal A_x$ is constant with value \begin{equation*} g\cdot |\{1\leq i\leq 2n\mathrel{|} \forall \sigma\in\mathfrak S(w_\sigma(i)=i\wedge \lambda_\sigma(i)=0)\}|. \end{equation*} \end{thm} \begin{remark} For $F=\QQ$, we recover the result \cite[Th\'eor\`eme 4.1]{genestier_ngo} of Ng\^o and Genestier. \end{remark} \begin{proof}[Proof of Theorem \ref{ThmPrank}] Let $1\leq i\leq 2n$ and $x\in\mathrm{Perm}$. Write $x=wu^\lambda$ with $w\in W,\lambda\in X$. By Propositions \ref{ProppRankFirstVersionGen} and \ref{PropEquivCondGen} it suffices to show the following equivalence. \begin{equation*} \widetilde{\Lambda}_i=x(\widetilde{\Lambda}_i)+\widetilde{\Lambda}_{i-1} \Leftrightarrow (w(i)=i\wedge \lambda(i)=0). \end{equation*} Consider the subset $\mathcal S=\{u^k e_j\mathrel{|} k\in\ZZ,\ 1\leq j\leq 2n\}$ of $\mathbb{F}(\!(u)\!)^{2n}$. Then $x$ induces a permutation of $\mathcal S$, namely $x(u^k e_j)=u^{\lambda(j)+k} e_{w(j)}$. We have $\widetilde{\Lambda}_{i}\cap \mathcal S=\widetilde{\Lambda}_{i-1}\cap \mathcal S\amalg \{u^{-1}e_i\}$, and $x\in\mathrm{Perm}$ implies $x(\widetilde{\Lambda}_{i-1}\cap \mathcal S)\subset \widetilde{\Lambda}_{i-1}\cap \mathcal S$. Consequently $u^{-1}e_i\in x(\widetilde{\Lambda}_{i}\cap \mathcal S)$ if and only if $x(u^{-1}e_i)=u^{-1}e_i$, which in turn is equivalent to $w(i)=i\wedge \lambda(i)=0$, as desired. \end{proof} \subsection{The density of the ordinary locus}\label{SecDensOrd} Denote by $\leq$ and $\ell$ the partial order and the length function on $\widetilde{W}$ defined in \cite[\textsection 2.1]{gy1}, respectively. We extend $\leq$ and $\ell$ to $\prod_{\sigma\in\mathfrak S} \widetilde{W}$ by setting $(x_\sigma)_\sigma\leq (x'_\sigma)_\sigma:\Leftrightarrow (\forall \sigma\in\mathfrak S:\ x_\sigma\leq x'_\sigma)$ and $\ell((x_\sigma)_\sigma)=\sum_\sigma\ell(x_\sigma)$. \begin{lem}\label{LemPropsOfKR} Let $x\in \prod_{\sigma\in\mathfrak S}\mathrm{Perm}$. The smooth $\mathbb{F}$-variety $\mathcal A_{x}$ is equidimensional of dimension $\ell(x)$. Furthermore the closure $\overline{\mathcal A}_{x}$ of $\mathcal A_{x}$ in $\mathcal A_{\mathbb{F}}$ is given by \begin{equation}\label{EqClosure} \overline{\mathcal A}_{x}=\coprod_{\substack{y\leq x}} \mathcal A_{y}. \end{equation} \end{lem} \begin{proof} As in Remark \ref{RemarkKRasScheme} there is, \'etale locally on $\mathcal A_\mathbb{F}$, an \'etale morphism $\beta:\mathcal A_\mathbb{F}\to M^{\mathrm{loc}}_\mathbb{F}$ with $\mathcal A_{x}=\beta^{-1}(M^{\mathrm{loc}}_x)$. In Theorem \ref{ThmIndicesSym} we have identified $M^{\mathrm{loc}}_x$ with the Schubert cell $\mathcal C_x\subset \prod_{\sigma\in\mathfrak S}\mathrm{GSp}_{2n}(\mathbb{F}(\!(u)\!))/I_{\mathrm{GSp}}(\mathbb{F})$ corresponding to $x$. The statements therefore follow from well-known properties of Schubert cells once we know that all KR strata are non-empty. This is true in the Siegel case by Genestier's result \cite[Proposition 1.3.2]{genestier_semi}, in the case that $p$ is unramified in $F$ by the result \cite[Theorem 2.5.2(1)]{goren} of Goren and Kassaei, and in the ramified case by a yet to be published result of Yu \cite{fakeyu}. \end{proof} Our next goal is to generalize the result \cite[Corollaire 4.3]{genestier_ngo} of Ng\^o and Genestier on the density of the ordinary locus. Denote by $\mathcal A^{(ng)} \subset \mathcal A(\mathbb{F})$ the subset where the $p$-rank of the underlying abelian variety is equal to $ng$. By the determinant condition imposed in the definition of $\mathcal A$ this is precisely the ordinary locus in $\mathcal A(\mathbb{F})$. \begin{cor}\label{CorDense} The ordinary locus $\mathcal A^{(ng)}$ is dense in $\mathcal A(\mathbb{F})$ if and only if $p$ is totally ramified in $F$. \end{cor} \begin{proof} Let $\mu=(e^{(g)},0^{(g)})\in X$. Our subset $\mathrm{Perm}\subset \widetilde{W}$ is precisely the set denoted by $\mathrm{Perm}(\mu)$ in \cite{haines_ngo}. By \cite[Theorem 10.1]{haines_ngo}, we have \begin{equation}\label{EqPermAdm} \mathrm{Perm}(\mu)=\mathrm{Adm}(\mu), \end{equation} where $\mathrm{Adm}(\mu):=\{x\in \widetilde{W}\mathrel{|} \exists w\in W:\ x\leq u^{w(\mu)}\}$. Write $\mathfrak M=\{x\in \widetilde{W}\mathrel{|} \exists w\in W:\ x= u^{w(\mu)}\}$. Then \eqref{EqPermAdm} implies that $\prod_{\sigma\in\mathfrak S}\mathfrak M$ is precisely the subset of maximal elements for $\leq$ in $\prod_{\sigma\in\mathfrak S}\mathrm{Perm}$. Denote by $\Delta_{\mathfrak M}\subset \prod_{\sigma\in\mathfrak S}\mathfrak M$ the diagonal. By Theorem \ref{ThmPrank} we have $\mathcal A^{(ng)}=\coprod_{x\in\Delta_{\mathfrak M}}\mathcal A_x$. The statement therefore follows from \eqref{EqClosure} by noting that $\Delta_\mathfrak M=\prod_{\sigma\in\mathfrak S}\mathfrak M$ if and only if $p$ is totally ramified in $F$. \end{proof} \subsection{An explicit example: Hilbert-Blumenthal modular varieties}\label{SecHilbBlum} In this section we use the explicit case of the Hilbert-Blumenthal modular varieties to illustrate how Theorem \ref{ThmPrank} and the KR stratification in general yield results about the geometry of the moduli spaces $\mathcal A$. We also compare these results to some of those obtained by Stamm in \cite{stamm}. Assume from now on that $p$ is \emph{inert} in $\mathcal O_F$, so that we have $e=1$ and $f=g$. Assume also that $\dim_F V=2$, so that $n=1$. Let us start with a discussion of the index set $\prod_{\sigma\in \mathfrak S} \mathrm{Perm}$ of the KR stratification. From Definition \ref{DefnPermSym} one immediately obtains that the subset $\mathrm{Perm}\subset\widetilde{W}$ is given by $\mathrm{Perm}=\{u^{(1,0)},u^{(0,1)},(1,2)u^{(1,0)}\}$. To put this set into a group theoretic perspective, we recall the setup described in \cite[\textsection 2.1]{gy1} in this easy special case. Consider the elements $\tau=(1,2)u^{(1,0)}, s_1=(1,2)$ and $s_0=(1,2)u^{(1,-1)}$ of $\widetilde{W}$. The subgroup $W_a$ of $\widetilde{W}$ generated by $s_0$ and $s_1$ is a Coxeter group on the generators $s_0$ and $s_1$, and we denote by $\leq$ and $\ell$ the corresponding Bruhat order and length function on $W_a$, respectively. Denoting by $\Omega$ the cyclic subgroup of $\widetilde{W}$ generated by $\tau$, we have $\widetilde{W}=W_a\rtimes \Omega$. The extension of $\leq$ and $\ell$ to $\widetilde{W}$ is given by $w'\tau'\leq w''\tau'':\Leftrightarrow (w'\leq w''\wedge \tau'=\tau'')$ and $\ell(w'\tau'):=\ell(w')$, for $w',w''\in W_a$ and $\tau',\tau''\in \Omega$. We extend $\leq$ and $\ell$ to $\prod_{\sigma\in\mathfrak S} \widetilde{W}$ as in Section \ref{SecDensOrd}. We see that \begin{equation*} \mathrm{Perm}=\{s_1\tau, s_0\tau, \tau\}\subset W_a\tau. \end{equation*} The Bruhat order on $\mathrm{Perm}$ is determined by the non-trivial relations $\tau\leq s_1\tau$ and $\tau\leq s_0\tau$, while the length function on $\mathrm{Perm}$ is given by $\ell(\tau)=0$ and $\ell(s_1\tau)=\ell(s_0\tau)=1$. Let us state Theorem \ref{ThmPrank} in this special case. Denote by $\mathcal A^{(0)} \subset \mathcal A(\mathbb{F})$ and $\mathcal A^{(g)} \subset \mathcal A(\mathbb{F})$ the subsets where the $p$-rank of the underlying abelian variety is equal to $0$ and $g$, respectively. \begin{prop}\label{PropPrankHilbert} We have \begin{equation*} \mathcal A(\mathbb{F})=\mathcal A^{(0)} \amalg \mathcal A^{(g)}. \end{equation*} The ordinary locus $\mathcal A^{(g)}$ is the union of only two KR strata, namely those corresponding to the elements $((s_1\tau)^{(g)})=(s_1\tau,s_1\tau,\dots,s_1\tau)$ and $((s_0\tau)^{(g)})=(s_0\tau,s_0\tau,\dots,s_0\tau)$ of $\prod_{\sigma\in\mathfrak S}\mathrm{Perm}$. The $p$-rank on all other KR strata is equal to 0. \end{prop} \begin{lem} The maximal elements in $\prod_{\sigma\in\mathfrak S}\mathrm{Perm}$ for the Bruhat order are precisely the elements of length $2^g$ in $\prod_{\sigma\in\mathfrak S}\mathrm{Perm}$. The set of these maximal elements is given by $\prod_{\sigma\in\mathfrak S}\{s_1\tau,s_0\tau\}$.\qed \end{lem} From the preceding results, we obtain without any additional work the following theorem. \begin{thm}\label{TheoremHB} Let $g\geq 2$. Then \begin{equation*} \mathcal A_{\mathbb{F}}=\overline{\mathcal A}_{((s_1\tau)^{(g)})}\cup \overline{\mathcal A}_{((s_0\tau)^{(g)})}\cup \mathcal A^{(0)}. \end{equation*} Each of $\overline{\mathcal A}_{((s_1\tau)^{(g)})}, \overline{\mathcal A}_{((s_0\tau)^{(g)})}$ and $\mathcal A^{(0)}$ is equidimensional of dimension $2^g$, and hence so is $\mathcal A_{\mathbb{F}}$. More precisely, $\mathcal A^{(0)}$ is the union \begin{equation*} \mathcal A^{(0)}=\bigcup_{\substack{x\in \prod_{\sigma\in\mathfrak S}\{s_1\tau,s_0\tau\}\\ x\neq ((s_1\tau)^{(g)}),((s_0\tau)^{(g)})}} \overline{\mathcal A}_{x} \end{equation*} of $2^g-2$ closed subsets, all equidimensional of dimension $2^g$. Furthermore, we have \begin{equation*} \overline{\mathcal A}_{((s_1\tau)^{(g)})}\cap \overline{\mathcal A}_{((s_0\tau)^{(g)})}\subset \mathcal A^{(0)}. \end{equation*} \end{thm} Taking $g=2$, we recover \cite[Theorem 2 (p.\ 408)]{stamm}. Note that for $g=2$, the set $\mathcal A^{(0)}$ is precisely the supersingular locus in $\mathcal A_{\mathbb{F}}$, because a 2-dimensional abelian variety is supersingular if and only if its $p$-rank is equal to zero. \section{The unitary PEL datum}\label{SecPELUni} Let $n\in\NN_{\geq 1}$. In Sections \ref{SecUni} through \ref{SecUni3} we will be concerned with the PEL datum consisting of the following objects. \begin{enumerate} \item An imaginary quadratic extension $F/F_0$ of a totally real extension $F_0/\QQ$. Let $g_0=[F_0:\QQ]$ and $g=[F:\QQ]$, so that $g=2g_0$. \item The non-trivial element $\ast$ of $\Gal(F/F_0)$. \item An $n$-dimensional $F$-vector space $V$. \item The symplectic form $\left(\cdot,\cdot\right):V\times V\to \QQ$ on the underlying $\QQ$-vector space of $V$ constructed as follows: Fix once and for all a $\ast$-skew-hermitian form $\left(\cdot,\cdot\right)':V\times V\to F$ (i.e.\ $\pairt{av}{bw}'=ab^\ast\pairt{v}{w}'$ and $\pairt{v}{w}'=-\pairt{w}{v}'^\ast$ for $v,w\in V,\ a,b\in F$). Define $\left(\cdot,\cdot\right)=\tr_{F/\QQ}\circ \left(\cdot,\cdot\right)'$. \item The element $J\in \End_{B\otimes \RR}(V\otimes \RR)$ to be defined separately in each case, see Sections \ref{SecPELUni1}, \ref{SecPELUni2} and \ref{SecPELUni3}. \end{enumerate} \begin{remark}\label{RemGUni1} Denote by $\mathrm{GU}_{\left(\cdot,\cdot\right)'}$ the $F_0$-group given on $R$-valued points by $\mathrm{GU}_{\left(\cdot,\cdot\right)'}(R)=\{g\in \mathrm{GL}_{F\otimes_{F_0} R}(V\otimes_{F_0} R)\mathrel{|} \exists c=c(g)\in R^\times \forall x,y\in V\otimes_{F_0} R:\ \pairt{gx}{gy}'_{R}=c\pairt{x}{y}'_{R} \}$. Then the reductive $\QQ$-group $G$ associated with the above PEL datum fits into the following cartesian diagram. \begin{equation*} \xymatrix{ G\ar[d]_c \,\ar@{^{(}->}[r] & \Res_{F_0/\QQ} \mathrm{GU}_{\left(\cdot,\cdot\right)'}\ar[d]^c\\ \GG_{m,\QQ}\,\ar@{^{(}->}[r]& \Res_{F_0/\QQ}\GG_{m,F_0}. } \end{equation*} \end{remark} \section{The ramified unitary case}\label{SecUni} \subsection{The PEL datum}\label{SecPELUni1} We start with the PEL datum defined in Section \ref{SecPELUni}. We assume that $p\mathcal O_{F_0}=(\mathcal P_0)^{e_0}$ for a single prime $\mathcal P_0$ of $\mathcal O_{F_0}$ and that $\mathcal P_0\mathcal O_F=\mathcal P^2$ for a prime $\mathcal P$ of $\mathcal O_F$. Write $e=2e_0$, so that $p\mathcal O_F=\mathcal P^e$. Denote by $f=[k_{\mathcal P_0}:\mathbb{F}_p]$ the corresponding inertia degree, so that $g=ef$ and $g_0=fe_0$. Fix once and for all uniformizers $\pi_0$ of $\mathcal O_{F_0}\otimes\ZZ_{(p)}$ and $\pi$ of $\mathcal O_{F}\otimes\ZZ_{(p)}$, satisfying $\pi^2=\pi_0$. We have $\pi^\ast=-\pi$. For typographical reasons, we denote the ring of integers in $F_\mathcal P$ by $\mathcal O_{\mathcal P}$ and the ring of integers in $(F_0)_{\mathcal P_0}$ by $\mathcal O_{\mathcal P_0}$. Denote by $\mathfrak C=\mathfrak C_{\mathcal O_{\mathcal P}\mathrel{|} \ZZ_p},\ \mathfrak C_0=\mathfrak C_{\mathcal O_{\mathcal P_0}\mathrel{|} \ZZ_p}$ and $\mathfrak C'=\mathfrak C_{\mathcal O_{\mathcal P}\mathrel{|} \mathcal O_{\mathcal P_0}}$ the corresponding inverse differents. Then $\mathfrak C_0=(\pi_0^{-k})$ for some $k\in\NN$. The extension $F_\mc P/(F_0)_{\mc P_0}$ is tamely ramified, so that $\mathfrak C'=(\pi^{-1})$. The equality $\mathfrak C=\mathfrak C'\cdot \mathfrak C_0$ then implies that $\mathfrak C=(\pi^{-2k-1})$ and we denote by $\delta=\pi^{-2k-1}$ the corresponding generator of $\mathfrak C$. It satisfies $\delta^\ast=-\delta$. Consequently the form $\delta^{-1}\left(\cdot,\cdot\right)'_{\QQ_p}:V_{\QQ_p}\times V_{\QQ_p}\to F_\mc P$ is $\ast$-hermitian and we assume that it \emph{splits}, i.e.\ that there is a basis $(e_1,\dots,e_n)$ of $V_{\QQ_p}$ over $F_\mc P$ such that $\pairt{e_i}{e_{n+1-j}}_{\QQ_p}'=\delta\delta_{ij}$ for $1\leq i,j\leq n$. Let $0\leq i<n$. We denote by $\Lambda_i$ the $\mathcal O_{\mathcal P}$-lattice in $V_{\QQ_p}$ with basis \begin{equation*} \mathfrak E_i=(\pi^{-1}e_1,\ldots,\pi^{-1}e_i,e_{i+1},\ldots,e_{n}). \end{equation*} For $k\in\ZZ$ we further define $\Lambda_{nk+i}=\pi^{-k}\Lambda_i$ and we denote by $\mathfrak E_{nk+i}$ the corresponding basis obtained from $\mathfrak E_i$. Then $\mathcal L=(\Lambda_i)_i$ is a complete chain of $\mathcal O_{\mathcal P}$-lattices in $V_{\QQ_p}$. For $i\in\ZZ$, the dual lattice $\Lambda_i^\vee:=\{x\in V_{\QQ_p}\mathrel{|} \pairt{x}{\Lambda_i}_{\QQ_p}\subset\ZZ_p\}$ of $\Lambda_i$ is given by $\Lambda_{-i}$. Consequently the chain $\mathcal L$ is self-dual. Let $i\in\ZZ$. We denote by $\rho_i:\Lambda_{i}\to \Lambda_{i+1}$ the inclusion, by $\vartheta_i:\Lambda_{n+i}\to \Lambda_i$ the isomorphism given by multiplication with $\pi$ and by $\left(\cdot,\cdot\right)_i:\Lambda_i\times \Lambda_{-i}\to \ZZ_p$ the restriction of $\left(\cdot,\cdot\right)_{\QQ_p}$. Then $(\Lambda_{i},\rho_{i},\vartheta_{i},\left(\cdot,\cdot\right)_{i})_i$ is a polarized chain of $\mathcal O_{F_\mc P}$-modules of type $(\mathcal L)$, which, by abuse of notation, we also denote by $\mathcal L$. Denote by $\left<\cdot,\cdot\right>_i:\Lambda_i\times \Lambda_{-i}\to \mathcal O_{\mathcal P}$ the restriction of the $\ast$-hermitian form $\delta^{-1}\left(\cdot,\cdot\right)'_{\QQ_p}$, and by $H_i$ the matrix describing $\left<\cdot,\cdot\right>_i$ with respect to $\mathfrak E_i$ and $\mathfrak E_{-i}$. We have \begin{equation} \label{EqDescribingMatrixUni} H_i=\text{anti-diag}((-1)^{a_{i,1}},\dots,(-1)^{a_{i,n}}) \end{equation} for some $a_{i,1},\dots,a_{i,n}\in\ZZ/2\ZZ$. Denote by $\Sigma_0$ the set of all embeddings $F_0\hookrightarrow\RR$ and by $\Sigma$ the set of all embeddings $F\hookrightarrow \CC$. For each $\sigma\in\Sigma_0$, we denote by $\tau_{\sigma,1},\tau_{\sigma,2}\in\Sigma$ the two embeddings with $\tau_{\sigma,j}\big|_{F_0}=\sigma$. Of course we have $\tau_{\sigma,2}=\tau_{\sigma,1}\circ \ast$. We obtain isomorphisms \begin{align} \label{FRdecomp} F\otimes_\QQ \RR &= \prod_{\sigma\in\Sigma_0}\CC, &F\ni x&\mapsto (\tau_{\sigma,1}(x))_\sigma,\\ \label{FCdecomp} F\otimes_\QQ \CC &= \prod_{\sigma\in\Sigma_0}\CC\times\CC,&F\ni x&\mapsto (\tau_{\sigma,1}(x),\tau_{\sigma,2}(x))_\sigma \end{align} of $\RR$- and $\CC$-algebras, respectively. The isomorphism \eqref{FRdecomp} induces a decomposition $V\otimes\RR = \prod_{\sigma\in\Sigma_0} V_\sigma$ into $\CC$-vector spaces $V_\sigma$ and $\left(\cdot,\cdot\right)'_\RR$ decomposes into the product of skew-hermitian forms $\left(\cdot,\cdot\right)'_\sigma:V_\sigma\times V_\sigma\to \CC,\ \sigma\in \Sigma_0$. For each $\sigma\in \Sigma_0$, there are $r_\sigma,s_\sigma\in\NN$ with $r_\sigma+s_\sigma=n$ and a basis $\mathfrak B_\sigma$ of $V_\sigma$ over $\CC$ such that $\left(\cdot,\cdot\right)'_\sigma$ is described by the matrix $D_\sigma=\diag(i^{(r_\sigma}),(-i)^{(s_\sigma)})$ with respect to $\mathfrak B_\sigma$. Denote by $J_\sigma$ the endomorphism of $V_\sigma$ described by the matrix $D_\sigma$ with respect to $\mathfrak B_\sigma$. We complete the description of the PEL datum by defining $J:=\prod_{\sigma\in \Sigma_0}J_\sigma\in \End_{B\otimes \RR}(V\otimes \RR)$. \subsection{The determinant morphism}\label{SecDetMorUni} The isomorphism \eqref{FCdecomp} induces a decomposition $V\otimes\CC = \prod_{\sigma\in\Sigma_0} (V_{\tau_{\sigma,1}}\times V_{\tau_{\sigma,2}})$ into $\CC$-vector spaces $V_{\tau_{\sigma,j}}$. The basis $\mathfrak B_\sigma$ of $V_\sigma$ induces bases $\mathfrak B_{\tau_{\sigma,j}}$ of $V_{\tau_{\sigma,j}}$ over $\CC$, and the endomorphism $J_{\sigma,\CC}$ decomposes into the product of endomorphisms $J_{\tau_{\sigma,j}}$ of $V_{\tau_{\sigma,j}}$. We find that $J_{\tau_{\sigma,1}}$ is described by the matrix $D_\sigma$ with respect to $\mathfrak B_{\tau_{\sigma,1}}$, while $J_{\tau_{\sigma,2}}$ is described by the matrix $-D_\sigma$ with respect to $\mathfrak B_{\tau_{\sigma,2}}$. Denote by $V_{-i}$ the $(-i)$-eigenspace of $J_\CC$. From the explicit description of the $J_{\tau_{\sigma,j}}$ with respect to the $\mathfrak B_{\tau_{\sigma,j}}$, one concludes that $V_{-i}$ is the $\mathcal O_F\otimes\CC$-module corresponding to the $\prod_{\sigma\in\Sigma_0}\CC\times\CC$-module $\prod_{\sigma\in\Sigma_0}\CC^{s_\sigma}\times \CC^{r_\sigma}$ under \eqref{FCdecomp}. Let $E'$ be the Galois closure of $F$ inside $\CC$ and choose a prime $\mathcal Q'$ of $E'$ over $\mathcal P$. In absolute analogy to \eqref{FCdecomp}, we have a decomposition \begin{equation}\label{EqFEDecomp} F\otimes_\QQ E'=\prod_{\sigma\in \Sigma_0}E'\times E'. \end{equation} Let $M$ be the $\mathcal O_F\otimes E'$-module corresponding to the $\prod_{\sigma\in \Sigma_0}E'\times E'$-module $\prod_{\sigma\in\Sigma_0}(E')^{s_\sigma}\times (E')^{r_\sigma}$ under \eqref{EqFEDecomp}. From the present discussion we obtain an identification $M\otimes_{E'}\CC=V_{-i}$ of $\mathcal O_F\otimes\CC$-modules. Let $\mathfrak B$ be a basis of $M$ over $E'$ and denote by $M_0$ the $\mathcal O_F\otimes \mathcal O_{E'}$-module generated by $\mathfrak B$. Then $M_0$ is an $\mathcal O_F\otimes \mathcal O_{E'}$-stable $\mathcal O_{E'}$-lattice $M_0$ in $M$. In particular, the morphism $\det_{V_{-i}}:V_{\mathcal O_F\otimes \CC}\to \mathbb{A}^1_\CC$ descends to the morphism $\det_{M_0}:V_{\mathcal O_F\otimes \mathcal O_{E'}}\to \mathbb{A}^1_{\mathcal O_{E'}}$. \subsection{The special fiber of the determinant morphism} We write $\mathfrak S=\Gal(k_{\mathcal P}/\mathbb{F}_p)=\Gal(k_{\mathcal P_0}/\mathbb{F}_p)$. We fix once and for all an embedding $\iota_{\mathcal Q'}:k_{\mathcal Q'}\hookrightarrow \mathbb{F}$. We consider $\mathbb{F}$ as an $\mathcal O_{E'}$-algebra via the composition $\mathcal O_{E'}\xrightarrow{\rho_{\mathcal Q'}}k_{\mathcal Q'}\overset{\iota_{\mathcal Q'}}{\hookrightarrow} \mathbb{F}$. Also $\iota_{\mathcal Q'}$ induces an embedding $\iota_{\mathcal P}:k_{\mathcal \mathcal P}\hookrightarrow \mathbb{F}$ and thereby an identification of the set of all embeddings $k_{\mathcal P}\hookrightarrow \mathbb{F}$ with $\mathfrak S$. Our choice of uniformizer $\pi$ induces a canonical isomorphism \begin{align} \label{decompspecialfiberfirstUni} \mathcal O_{F}\otimes\mathbb{F}&=\prod_{\sigma\in\mathfrak S} \mathbb{F}[u]/(u^e). \end{align} \begin{prop}\label{PropDeterminantSpecialFiberUni} Let $x\in \mathcal O_F$ and let $(p_\sigma)_\sigma\in \prod_{\sigma\in\mathfrak S} \mathbb{F}[u]/(u^e)$ be the element corresponding to $x\otimes 1$ under \eqref{decompspecialfiberfirstUni}. Then \begin{equation*} \chi_\mathbb{F}(x| M_0\otimes_{\mathcal O_{E'}}\mathbb{F})=\prod_{\sigma\in\mathfrak S}\bigl(T-p_\sigma(0)\bigr)^{ne_0} \end{equation*} in $\mathbb{F}[T]$. \end{prop} \begin{proof} Reduce $\chi_{\mathcal O_{E'}}(x|M_0)$ modulo $\mathcal Q'$, using Lemma \ref{LemResidueMapinNonGaloisCase}. \end{proof} Denote by $E=\QQ(\tr_\CC(x\otimes 1|V_{-i});\ x\in F)$ the reflex field and define $\mathcal Q=\mathcal Q'\cap \mathcal O_E$. The morphism $\det_{V_{-i}}$ is defined over $\mathcal O_E$. \subsection{The local model} For the chosen PEL datum, Definition \ref{DefnLocalModelGen} amounts to the following. \begin{defn}\label{DefnLocalModelUni} The local model $M^{\mathrm{loc}}$ is the functor on the category of $\mathcal O_{E_\mathcal Q}$-algebras with $M^{\mathrm{loc}}(R)$ the set of tuples $(t_i)_{i\in \ZZ}$ of $\mathcal O_{F}\otimes R$-submodules $t_i\subset \Lambda_{i,R}$ satisfying the following conditions for all $i\in\ZZ$. \renewcommand\theenumi {\alph{enumi}} \begin{enumerate} \item\label{DefnLocalModelUni-Functoriality} $\rho_{i,R}(t_i)\subset t_{i+1}$. \item\label{DefnLocalModelUni-Projectivity} The quotient $\Lambda_{i,R}/t_i$ is a finite locally free $R$-module. \item\label{DefnLocalModelUni-Determinant} We have an equality \begin{equation*} \det_{\Lambda_{i,R}/t_i}=\det_{V_{-i}}\otimes_{\mathcal O_E}R \end{equation*} of morphisms $V_{\mathcal O_{F}\otimes R}\to \mathbb{A}^1_R$. \item\label{DefnLocalModelUni-DualityCondition} Under the pairing $\left(\cdot,\cdot\right)_{i,R}:\Lambda_{i,R}\times \Lambda_{-i,R}\to R$, the submodules $t_i$ and $t_{-i}$ pair to zero. \item\label{DefnLocalModelUni-Periodicity} $\vartheta_i(t_{n+i})=t_i$. \end{enumerate} \end{defn} \subsection{The special fiber of the local model} For $i\in\ZZ$, denote by $\overline{\Lambda}_i$ the free $\mathbb{F}[u]/u^e$-module $(\mathbb{F}[u]/u^e)^{n}$ and by $\overline{\mathfrak E}_i$ its canonical basis. Consider the $\mathbb{F}$-automorphism $\overline{\ast}:\mathbb{F}[u]/u^e\to \mathbb{F}[u]/u^e,\ u\mapsto -u$. Denote by $\overline{\left<\cdot,\cdot\right>}_i:\overline{\Lambda}_i\times \overline{\Lambda}_{-i}\to \mathbb{F}[u]/u^e$ the $\overline{\ast}$-sesquilinear form described by the matrix $H_i$ of \eqref{EqDescribingMatrixUni} with respect to $\overline{\mathfrak E}_i$ and $\overline{\mathfrak E}_{-i}$. Denote by $\overline{\vartheta}_i:\overline{\Lambda}_{n+i}\to \overline{\Lambda}_i$ the identity morphism. For $k\in\ZZ$ and $0\leq i<n$, let $\overline{\rho}_{nk+i}:\overline{\Lambda}_{nk+i}\to \overline{\Lambda}_{nk+i+1}$ be the morphism described by the matrix $\mathrm{diag}(1^{(i)},u,1^{(n-i-1)})$ with respect to $\overline{\mathfrak E}_{nk+i}$ and $\overline{\mathfrak E}_{nk+i+1}$. \begin{defn}\label{DefnSpecialLocalModelUni} Define a functor $M^{e,n}$ on the category of $\mathbb{F}$-algebras with $M^{e,n}(R)$ the set of tuples $(t_i)_{i\in \ZZ}$ of $R[u]/u^e$-submodules $t_i\subset \overline{\Lambda}_{i,R}$ satisfying the following conditions for all $i\in\ZZ$. \renewcommand\theenumi {\alph{enumi}} \begin{enumerate} \item $\overline{\rho}_{i,R}(t_i)\subset t_{i+1}$. \item The quotient $\overline{\Lambda}_{i,R}/t_i$ is finite locally free over $R$. \item\label{DefnSpecialLocalModelUni-Determinant} For all $p\in R[u]/u^e$, we have \begin{equation*} \chi_R(p|\overline{\Lambda}_{i,R}/t_i)=\bigl(T-p(0)\bigr)^{ne_0} \end{equation*} in $R[T]$. \item $t_i^{\perp,\overline{\left<\cdot,\cdot\right>}_{i,R}}=t_{-i}$. \item $\overline{\vartheta}_i(t_{n+i})=t_i$. \end{enumerate} \end{defn} Let $i\in\ZZ$. From \eqref{decompspecialfiberfirstUni} we obtain an isomorphism \begin{equation}\label{LambdaInSpecialFiberUni} \Lambda_{i,\mathbb{F}}=\prod_{\sigma\in\mathfrak S}\overline{\Lambda}_{i} \end{equation} by identifying the basis $\mathfrak E_{i,\mathbb{F}}$ with the product of the bases $\overline{\mathfrak E}_i$. Under this identification, the morphism $\rho_{i,\mathbb{F}}$ decomposes into the morphisms $\overline{\rho}_i$, the pairing $\left<\cdot,\cdot\right>_{i,\mathbb{F}}$ decomposes into the pairings $\overline{\left<\cdot,\cdot\right>}_i$ and the morphism $\vartheta_{i,\mathbb{F}}$ decomposes into the morphisms $\overline{\vartheta}_i$. Let $R$ be an $\mathbb{F}$-algebra and let $(t_i)_{i\in\ZZ}$ be a tuple of $\mathcal O_F\otimes R$-submodules $t_i\subset \Lambda_{i,R}$. Then \eqref{LambdaInSpecialFiberUni} induces decompositions $t_i=\prod_{\sigma\in\mathfrak S} t_{i,\sigma}$ into $R[u]/u^e$-submodules $t_{i,\sigma}\subset\overline{\Lambda}_{i,R}$. The following statement is then clear (cf.\ the proof of Proposition \ref{PropDecompositionofLocalModel}). \begin{prop}\label{PropDecompositionofLocalModelUni} The morphism $M^{\mathrm{loc}}_\mathbb{F}\to \prod_{\sigma\in\mathfrak S} M^{e,n}$ given on $R$-valued points by \begin{equation} \label{EqDecompLocModelUni} \begin{aligned} M^{\mathrm{loc}}_\mathbb{F}(R)&\to \prod_{\sigma\in\mathfrak S} M^{e,n}(R),\\ (t_i)&\mapsto \left((t_{i,\sigma})_i\right)_\sigma \end{aligned} \end{equation} is an isomorphism of functors on the category of $\mathbb{F}$-algebras. \end{prop} \subsection{The affine flag variety} This section deals with the affine flag variety for the ramified unitary group. Our discussion is based on and has greatly profited from \cite{pr}, \cite{pr3} and \cite{smithling_unitary_odd}, \cite{smithling_unitary_even}. Let $R$ be an $\mathbb{F}$-algebra. Consider the extension $R[\![u]\!]/R[\![u_0]\!]$ with $u_0=u^2$. Also consider the $R(\!(u_0)\!)$-automorphism $\widetilde{\ast}:R(\!(u)\!)\to R(\!(u)\!),\ u\mapsto -u$. Let $\widetilde{\left<\cdot,\cdot\right>}$ be the $\widetilde{\ast}$-hermitian form on $R(\!(u)\!)^{n}$ described by the matrix $\widetilde{I}_n$ with respect to the standard basis of $R(\!(u)\!)^{n}$ over $R(\!(u)\!)$. For a lattice $\Lambda$ in $R(\!(u)\!)^n$ we define $\Lambda^\vee:=\{x\in R(\!(u)\!)^{n}\mathrel{|} \widetilde{\pair{x}{\Lambda}}\subset R[\![u]\!]\}$. Recall from Section \ref{SecLattices} the standard lattice chain $\widetilde{\mathcal L}=(\widetilde{\Lambda}_i)_i$ in $R(\!(u)\!)^{n}$. Note that $(\widetilde{\Lambda}_i)^\vee=\widetilde{\Lambda}_{-i}$ for all $i\in\ZZ$. We denote by $\widetilde{\left<\cdot,\cdot\right>}_i:\widetilde{\Lambda}_i\times \widetilde{\Lambda}_{-i}\to R[\![u]\!]$ the restriction of $\widetilde{\left<\cdot,\cdot\right>}$. It is the perfect $\widetilde{\ast}$-sesquilinear pairing described by the matrix $H_i$ of \eqref{EqDescribingMatrixUni} with respect to $\widetilde{\mathfrak E}_i$ and $\widetilde{\mathfrak E}_{-i}$. In complete analogy with \cite[Definition A.41]{rz}, we have for an $\mathbb{F}[\![u_0]\!]$-algebra $R$ the notion of a polarized chain $\mathcal M=(M_i,\varrho_i:M_i\to M_{i+1},\theta_i:M_{n+i}\xrightarrow{\sim} M_i,\mathcal E_i:M_i\times M_{-i}\to\mathbb{F}[\![u]\!]\otimes_{\mathbb{F}[\![u_0]\!]} R)_{i\in\ZZ}$ of $\mathbb{F}[\![u]\!]\otimes_{\mathbb{F}[\![u_0]\!]} R$-modules of type $(\widetilde{\mathcal L})$ (cf.\ \cite[Definition 6.6.1]{diss}). The proof of \cite[Proposition A.43]{rz} then carries over without any changes to show the following result. \begin{prop}\label{PropRZbigUni} Let $R$ be an $\mathbb{F}[\![u_0]\!]$-algebra such that the image of $u_0$ in $R$ is nilpotent. Then any two polarized chains $\mathcal M,\mathcal N$ of $\mathbb{F}[\![u]\!]\otimes_{\mathbb{F}[\![u_0]\!]} R$-modules of type $(\widetilde{\mathcal L})$ are isomorphic locally for the \'etale topology on $R$. Furthermore the functor $\Isom(\mathcal M,\mathcal N)$ is representable by a smooth affine scheme over $R$. \end{prop} \begin{prop}\label{PropLiftingIsosUni} Let $R$ be an $\mathbb{F}$-algebra and let $\mathcal M,\mathcal N$ be polarized chains of $\mathbb{F}[\![u]\!]\otimes_{\mathbb{F}[\![u_0]\!]} R[\![u_0]\!]$-modules of type $(\widetilde{\mathcal L})$. Then the canonical map $\Isom(\mathcal M,\mathcal N)(R[\![u_0]\!])\to \Isom(\mathcal M,\mathcal N)(R[\![u_0]\!]/u_0^m)$ is surjective for all $m\in\NN_{\geq 1}$. In particular $\mathcal M$ and $\mathcal N$ are isomorphic locally for the \'etale topology on $R$. \end{prop} \begin{proof} Analogous to the proof of Proposition \ref{PropLiftingIsosSym}. \end{proof} Denote by $\mathrm{U}=\mathrm{U}_n$ and $\mathrm{GU}=\mathrm{GU}_n$ the $\mathbb{F}(\!(u_0)\!)$-groups with \begin{equation*} \mathrm{U}(R)=\{g\in \mathrm{GL}_{n}(K\otimes_{K_0} R)\mathrel{|} g^t\widetilde{I}_n g^{\widetilde{\ast}}=\widetilde{I}_n\} \end{equation*} and \begin{align*} \mathrm{GU}(R)=\left\{g\in \mathrm{GL}_{n}(K\otimes_{K_0} R)\mathrel{|} \exists c=c(g)\in R^\times:\ g^t\widetilde{I}_n g^{\widetilde{\ast}}=c\widetilde{I}_n\right\}. \end{align*} \begin{defn}[{\cite[\textsection 4.2]{smithling_unitary_odd}, \cite[\textsection 6.2]{smithling_unitary_even}}] Let $R$ be an $\mathbb{F}$-algebra and let $(L_i)_i$ be a lattice chain in $R(\!(u)\!)^{n}$. \begin{enumerate} \item Let $r\in\ZZ$. The chain $(L_i)_i$ is called \emph{$r$-self-dual} if \begin{equation*} \forall i\in\ZZ:\ L_i^\vee=u_0^r L_{-i}. \end{equation*} Denote by $\mathcal F^{(r)}_\mathrm{U}$ the functor on the category of $\mathbb{F}$-algebras with $\mathcal F_{\mathrm{U}}^{(r)}(R)$ the set of $r$-self-dual lattice chains in $R(\!(u)\!)^{n}$. \item The chain $(L_i)_i$ is called \emph{self-dual} if Zariski locally on $R$ there is an $a\in R(\!(u_0)\!)^\times$ such that \begin{equation}\label{EquSelfdualUni} \forall i\in\ZZ:\ L_i^\vee=a L_{-i}. \end{equation} Denote by $\mathcal F_{\mathrm{GU}}$ the functor on the category of $\mathbb{F}$-algebras with $\mathcal F_{\mathrm{GU}}(R)$ the set of self-dual lattice chains in $R(\!(u)\!)^{n}$. \end{enumerate} \end{defn} Note that $\widetilde{\mathcal L}\in \mathcal F_\mathrm{U}^{(0)}(R)$. \begin{remark} Let $R$ be a reduced $\mathbb{F}$-algebra such that $\Spec R$ is connected. Then \begin{equation*} \mathcal F_\mathrm{GU}(R)=\bigcup_{r\in\ZZ} \mathcal F_\mathrm{U}^{(r)}(R). \end{equation*} \end{remark} \begin{proof} This follows directly from Lemma \ref{LemUnitsinLaurent}. \end{proof} \begin{remark}\label{RemarkLatticesAreChainsUni} Let $R$ be an $\mathbb{F}$-algebra and let $(L_i)_i\in \mathcal F_{\mathrm{U}}^{(0)}(R)$. For $i\in\ZZ$ denote by $\varrho_i:L_i\to L_{i+1}$ the inclusion, by $\theta_i:L_{n+i}\to L_i$ the isomorphism given by multiplication with $u$ and by $\mathcal E_i:L_i\times L_{-i}\to R[\![u]\!]$ the restriction of $\widetilde{\left<\cdot,\cdot\right>}$. Then $(L_i,\varrho_i,\theta_i,\mathcal E_i)$ is a polarized chain of $\mathbb{F}[\![u]\!]\otimes_{\mathbb{F}[\![u_0]\!]}R[\![u_0]\!]$-modules of type $(\widetilde{\mathcal L})$. \end{remark} Note that for an $\mathbb{F}$-algebra $R$, the canonical maps \begin{align*} \mathbb{F}[\![u]\!]\otimes_{\mathbb{F}[\![u_0]\!]}R[\![u_0]\!]&\to R[\![u]\!],\\ \mathbb{F}(\!(u)\!)\otimes_{\mathbb{F}(\!(u_0)\!)}R(\!(u_0)\!)&\to R(\!(u)\!) \end{align*} are isomorphisms. Consequently we can consider $\Lf_{u_0} \mathrm{GU}$ and $\Lf_{u_0} \mathrm{U}$ as subfunctors of $\Lf_u \mathrm{GL}_n$. Recall from Remark \ref{RemarkActionOfGLOnF} the subfunctor $I\subset \Lf \mathrm{GL}_{n}$. We define a subfunctor $I_\mathrm{GU}$ of $\Lf_{u_0} \mathrm{GU}$ by $I_\mathrm{GU}=\Lf_{u_0} \mathrm{GU}\cap I$. \begin{prop} The natural action of $\Lf_u \mathrm{GL}_{n}$ on $\mathcal F$ (cf.\ Remark \ref{RemarkActionOfGLOnF}) restricts to an action of $\Lf_{u_0} \mathrm{GU}$ on $\mathcal F_\mathrm{GU}$. Consequently we obtain an injective map \begin{equation*} \begin{aligned} \Lf_{u_0} \mathrm{GU}(R)/I_\mathrm{GU}(R)&\xrightarrow{\phi(R)} \mathcal F_\mathrm{GU}(R),\\ g&\xmapsto{\hphantom{\phi(R)}} g\cdot \widetilde{\mathcal L} \end{aligned} \end{equation*} for each $\mathbb{F}$-algebra $R$. The morphism $\phi$ identifies $\mathcal F_\mathrm{GU}$ with both the \'etale and the fpqc sheafification of the presheaf $\Lf_{u_0} \mathrm{GU}/I_\mathrm{GU}$. \end{prop} \begin{proof} Let $R$ be an $\mathbb{F}$-algebra. We claim that every $a\in R(\!(u_0)\!)^\times$ lies in the image of the map $c:\mathrm{GU}(R(\!(u_0)\!))\to R(\!(u_0)\!)^\times$ \'etale locally on $R$. Assuming this, one can proceed exactly as in the proof of Proposition \ref{PropFuncDescofFlag}. To prove the claim, first note that for $b\in R(\!(u)\!)$, the matrix $bI_n\in \mathrm{GU}(R(\!(u_0)\!))$ satisfies $c(bI_n)=bb^{\widetilde{\ast}}$. Lemma \ref{LemUnitsinLaurent} implies that Zariski locally on $R$, the element $a$ is of the form $a=u_0^k\upsilon(1+n)$ for some $k\in\ZZ$, a unit $\upsilon\in R[\![u_0]\!]^\times$ and a nilpotent element $n\in R(\!(u_0)\!)$. Consequently it suffices to show that each of $u_0^k,\upsilon$ and $1+n$ is of the form $bb^{\widetilde{\ast}}$ for some $b\in R(\!(u)\!)$ \'etale locally on $R$. As $2\in R^\times$, one easily sees that $\upsilon$ is a square in $R[\![u_0]\!]^\times$ whenever $\upsilon(0)$ is a square in $R^\times$, which is the case \'etale locally on $R$. For $b=\sqrt{-1}u^k$, one has $bb^{\widetilde{\ast}}=u_0^k$. Finally, $1+n$ is a square in $R(\!(u_0)\!)^\times$; this follows from the Taylor expansion of $\sqrt{1+x}$ if one notes that $\binom{1/2}{l}\in \ZZ[\frac{1}{2}]$ for all $l\in\NN$. \end{proof} \subsection{Embedding the local model into the affine flag variety} Let $R$ be an $\mathbb{F}$-algebra. We consider an $R[u]/u^e$-module as an $R[\![u]\!]$-module via the canonical projection $R[\![u]\!]\to R[u]/u^e$. For $i\in\ZZ$ denote by $\alpha_i:\widetilde{\Lambda}_i\to \overline{\Lambda}_{i,R}$ the morphism described by the identity matrix with respect to $\widetilde{\mathfrak E}_i$ and $\overline{\mathfrak E}_i$. It induces an isomorphism $\widetilde{\Lambda}_i/u^e\widetilde{\Lambda}_i\xrightarrow{\sim}\overline{\Lambda}_{i,R}$. Clearly the following diagrams commute. \begin{equation*} \xymatrix{ \widetilde{\Lambda}_i\ar[d]_{\alpha_i}\ar@{}[r]|{\subset}&\widetilde{\Lambda}_{i+1}\ar[d]^{\alpha_{i+1}}\\ \overline{\Lambda}_{i,R}\ar[r]^-{\overline{\rho}_{i,R}}&\overline{\Lambda}_{i+1,R}, } \quad \xymatrix{ \widetilde{\Lambda}_i\times \widetilde{\Lambda}_{-i}\ar[r]^-{\widetilde{\left<\cdot,\cdot\right>}_i}\ar[d]_{\alpha_i\times \alpha_{-i}}& R[\![u]\!]\ar[d]\\ \overline{\Lambda}_{i,R}\times \overline{\Lambda}_{-i,R}\ar[r]^-{\overline{\left<\cdot,\cdot\right>}_{i,R}}& R[u]/u^e, } \quad \xymatrix{ \widetilde{\Lambda}_i\ar[d]_{\alpha_i}&\widetilde{\Lambda}_{n+i}\ar[d]^{\alpha_{n+i}}\ar[l]_{u \cdot}\\ \overline{\Lambda}_{i,R}&\overline{\Lambda}_{n+i,R}\ar[l]_{\overline{\vartheta}_{i,R}}. } \end{equation*} The following proposition allows us to consider $M^{e,n}$ as a subfunctor of $\mathcal F_\mathrm{U}^{(-e_0)}$. \begin{prop}[{\cite[\textsection 3.3]{pr3},\cite[\textsection 4.4-5.1]{smithling_unitary_odd}, \cite[\textsection 6.4-7.1]{smithling_unitary_even}}]\label{PropLocalModelintoFlagUni} There is an embedding $\alpha:M^{e,n}\hookrightarrow \mathcal F_\mathrm{U}^{(-e_0)}$ given on $R$-valued by \begin{align*} M^{e,n}(R)&\to \mathcal F_\mathrm{U}^{(-e_0)}(R),\\ (t_i)_i&\mapsto (\alpha_i^{-1}(t_i))_i. \end{align*} It induces a bijection from $M^{e,n}(R)$ onto the set of those $(L_i)_i\in \mathcal F_\mathrm{U}^{(-e_0)}(R)$ satisfying the following conditions for all $i\in\ZZ$. \begin{enumerate} \item $u^e\widetilde{\Lambda}_i\subset L_i\subset\widetilde{\Lambda}_i$. \item For all $p\in R[u]/u^e$, we have \begin{equation*} \chi_R(p|\widetilde{\Lambda}_i/L_i)=(T-p(0))^{ne_0} \end{equation*} in $R[T]$. Here $\widetilde{\Lambda}_i/L_i$ is considered as an $R[u]/u^e$-module using (1). \end{enumerate} \end{prop} \begin{proof} Identical to the proof of Proposition \ref{PropLocalModelintoFlag}. \end{proof} Note that $\overline{\mathcal L}=(\overline{\Lambda}_{i},\overline{\rho}_{i},\overline{\vartheta}_{i},\overline{\left<\cdot,\cdot\right>}_{i})_i$ is a polarized chain of $\mathbb{F}[\![u]\!]\otimes_{\mathbb{F}[\![u_0]\!]}\mathbb{F}[u_0]/u_0^{e_0}$-modules of type $(\widetilde{\mathcal L})$. In fact $\overline{\mathcal L}=\widetilde{\mathcal L}\otimes_{\mathbb{F}[\![u_0]\!]}\mathbb{F}[u_0]/u_0^{e_0}$. Let $R$ be an $\mathbb{F}$-algebra. There is an obvious action of $\Aut(\overline{\mathcal L})(R[u_0]/u_0^{e_0})$ on $M^{e,n}(R)$, given by $(\varphi_i)\cdot (t_{i})=(\varphi_i(t_i))$. The canonical morphism $R[\![u_0]\!]\to R[u_0]/u_0^{e_0}$ induces a morphism $\Aut(\widetilde{\mathcal L})(R[\![u_0]\!])\to \Aut(\overline{\mathcal L})(R[u_0]/u_0^{e_0})$ and we thereby extend this $\Aut(\overline{\mathcal L})(R[u_0]/u_0^{e_0})$-action on $M^{e,n}(R)$ to an $\Aut(\widetilde{\mathcal L})(R[\![u_0]\!])$-action. \begin{lem}\label{LemDifferentAutsSameOrbitsUni} Let $R$ be an $\mathbb{F}$-algebra and let $t\in M^{e,n}(R)$. We have $\Aut(\widetilde{\mathcal L})(R[\![u_0]\!])\cdot t=\Aut(\overline{\mathcal L})(R[u_0]/u_0^e)\cdot t$. \end{lem} \begin{proof} The map $\Aut(\widetilde{\mathcal L})(R[\![u_0]\!])\to \Aut(\overline{\mathcal L})(R[u_0]/u_0^{e_0})$ is surjective by Proposition \ref{PropLiftingIsosUni}. \end{proof} Define a subfunctor $I_\mathrm{U}$ of $\Lf_{u_0} \mathrm{U}$ by $I_{\mathrm{U}}=\Lf_{u_0} \mathrm{U}\cap I_\mathrm{GU}$. \begin{lem}\label{LemIvsI0Uni} We have $I_\mathrm{GU}(\mathbb{F})=\mathbb{F}[\![u_0]\!]^\times I_\mathrm{U}(\mathbb{F})$. \end{lem} \begin{proof} Analogous to the proof of Lemma \ref{LemIvsI0Sym}, noting that for $g\in I_\mathrm{GU}(\mathbb{F})$ one has $c(g)\in \mathbb{F}[\![u_0]\!]^\times$. \end{proof} \begin{lem}\label{LemIvsIprimeUni} Let $g\in I_\mathrm{U}(\mathbb{F})$. Then $g$ restricts to an automorphism $g_i:\widetilde{\Lambda}_i\xrightarrow{\sim}\widetilde{\Lambda}_i$ for each $i\in\ZZ$. The assignment $g\mapsto (g_i)_i$ defines an isomorphism $I_\mathrm{U}(\mathbb{F})\xrightarrow{\sim} \Aut(\widetilde{\mathcal L})(\mathbb{F}[\![u_0]\!])$. \end{lem} \begin{proof} Analogous to the proof of Lemma \ref{LemIvsIprime}. \end{proof} \begin{prop}\label{PropIndicesUni_2} Let $t\in M^{e,n}(\mathbb{F})$. Then $\alpha$ induces a bijection \begin{equation*}\label{EqOrbits1Uni} \Aut(\overline{\mathcal L})(\mathbb{F}[u_0]/u_0^{e_0})\cdot t\xrightarrow{\sim} I_{\mathrm{GU}}(\mathbb{F})\cdot \alpha(t). \end{equation*} Consequently we obtain an embedding \begin{equation*} \Aut(\overline{\mathcal L})(\mathbb{F}[u_0]/u_0^{e_0})\backslash M^{e,n}(\mathbb{F})\hookrightarrow I_{\mathrm{GU}}(\mathbb{F})\backslash \mathcal F_\mathrm{GU}(\mathbb{F}). \end{equation*} \end{prop} \begin{proof} Analogous to the proof of Proposition \ref{PropIndicesSym2}. \end{proof} Consider $\alpha':M^{e,n}(\mathbb{F})\hookrightarrow \mathcal F_\mathrm{GU}(\mathbb{F})\xrightarrow{\phi(\mathbb{F})^{-1}}\Lf_{u_0} {\mathrm{GU}}(\mathbb{F})/I_{\mathrm{GU}}(\mathbb{F})$. \begin{prop}\label{PropIndicesUni} Let $t\in M^{e,n}(\mathbb{F})$. Then $\alpha'$ induces a bijection \begin{equation*} \Aut(\overline{\mathcal L})(\mathbb{F}[u_0]/u_0^{e_0})\cdot t\xrightarrow{\sim} I_{\mathrm{GU}}(\mathbb{F})\cdot \alpha'(t). \end{equation*} Consequently we obtain an embedding \begin{equation*} \Aut(\overline{\mathcal L})(\mathbb{F}[u_0]/u_0^{e_0})\backslash M^{e,n}(\mathbb{F})\hookrightarrow I_{\mathrm{GU}}(\mathbb{F})\backslash \mathrm{GU}(\mathbb{F}(\!(u_0)\!))/I_{\mathrm{GU}}(\mathbb{F}). \end{equation*} \end{prop} \begin{proof} Clear from Proposition \ref{PropIndicesUni_2}, as the isomorphism $\phi(\mathbb{F})$ is in particular $I_{\mathrm{GU}}(\mathbb{F})$-equivariant. \end{proof} Let $R$ be an $\mathbb{F}$-algebra and $(\varphi_i)_i\in \Aut(\mathcal L)(R)$. The decomposition \eqref{LambdaInSpecialFiberUni} induces for each $i$ a decomposition of $\varphi_i:\Lambda_{i,R}\xrightarrow{\sim}\Lambda_{i,R}$ into the product of $R[u]/u^e$-linear automorphisms $\varphi_{i,\sigma}:\overline{\Lambda}_{i,R}\xrightarrow{\sim}\overline{\Lambda}_{i,R}$. The following statement is then clear (cf.\ the proof of Proposition \ref{PropDecompositionofLocalModel}). \begin{prop}\label{PropChainMorphismDecomposesUni} Let $R$ be an $\mathbb{F}$-algebra. The following map is an isomorphism, functorial in $R$. \begin{align*} \Aut(\mathcal L)(R)&\to \prod_{\sigma\in\mathfrak S} \Aut(\overline{\mathcal L})(R[u_0]/u_0^{e_0}),\\ (\varphi_i)_i&\mapsto ((\varphi_{i,\sigma})_i)_{\sigma\in\mathfrak S}. \end{align*} \end{prop} Consider the composition \begin{equation*} \tilde{\alpha}:M^{\mathrm{loc}}(\mathbb{F})\xrightarrow{\eqref{EqDecompLocModelUni}} \prod_{\sigma\in\mathfrak S} M^{e,n}(\mathbb{F})\xrightarrow{\prod_\sigma \alpha'} \prod_{\sigma\in\mathfrak S} \Lf_{u_0} {\mathrm{GU}}(\mathbb{F})/I_{\mathrm{GU}}(\mathbb{F}). \end{equation*} For $\sigma\in \mathfrak S$ denote by $\tilde{\alpha}_\sigma:M^{\mathrm{loc}}(\mathbb{F})\to \Lf_{u_0} {\mathrm{GU}}(\mathbb{F})/I_{\mathrm{GU}}(\mathbb{F})$ the corresponding component of $\tilde{\alpha}$. \begin{thm}\label{ThmIndicesUni} Let $t\in M^{\mathrm{loc}}(\mathbb{F})$. Then $\tilde{\alpha}$ induces a bijection \begin{equation*} \Aut(\mathcal L)(\mathbb{F})\cdot t\xrightarrow{\sim} \prod_{\sigma\in\mathfrak S} I_{\mathrm{GU}}(\mathbb{F})\cdot \tilde{\alpha}_\sigma(t). \end{equation*} Consequently we obtain an embedding \begin{equation*} \Aut(\mathcal L)(\mathbb{F})\backslash M^{\mathrm{loc}}(\mathbb{F})\hookrightarrow \prod_{\sigma\in\mathfrak S} I_{\mathrm{GU}}(\mathbb{F})\backslash \mathcal \mathrm{GU}(F(\!(u_0)\!))/I_{\mathrm{GU}}(\mathbb{F}). \end{equation*} \end{thm} \begin{proof} Identical to the proof of Theorem \ref{ThmIndicesSym}. \end{proof} \subsection{The extended affine Weyl group} As in \cite[3.2]{smithling_unitary_odd},\cite[3.2]{smithling_unitary_even}, we denote by $S$ the standard diagonal maximal split torus in $\mathrm{GU}$. Denote by $T$ the centralizer and by $N$ the normalizer of $S$ in $\mathrm{GU}$. By the discussion in \cite[3.4]{smithling_unitary_odd}, \cite[5.4]{smithling_unitary_even}, the Kottwitz homomorphism for $T$ is given by \begin{equation*} \kappa_T:T(\mathbb{F}(\!(u_0)\!))\to \ZZ^n,\quad \diag(x_1,\dots,x_n)\mapsto (\mathrm{val}_u(x_1),\dots,\mathrm{val}_u(x_n)). \end{equation*} Consequently the kernel $T\bigl(\mathbb{F}(\!(u_0)\!)\bigr)_1$ of $\kappa_T$ is equal to $T(\mathbb{F}(\!(u_0)\!))\cap D_n(\mathbb{F}[\![u]\!])$, with the intersection taking place in $\mathrm{GL}_n(\mathbb{F}(\!(u)\!))$. Here $D_n\subset \mathrm{GL}_n$ denotes the subgroup of diagonal matrices. By definition, the extended affine Weyl group of $\mathrm{GU}$ with respect to $S$ is given by $\widetilde{W}:=N(\mathbb{F}(\!(u_0)\!))/T(\mathbb{F}(\!(u_0)\!))_1$. Set \begin{equation*} W=\{w\in S_{n}\mathrel{|} \forall i\in\{1,\dots,n\}:\ w(i)+w(n+1-i)=n+1\} \end{equation*} and \begin{equation*} X=\{(x_1,\dots,x_n)\in\ZZ^n\mathrel{|} \exists r\in\ZZ\forall i\in\{1,\dots,n\}:\ x_i+x_{n+1-i}=2r\}. \end{equation*} We identify $W$ with a subgroup of $\mathrm{U}(\mathbb{F}(\!(u_0)\!))$ via $W\ni w\mapsto A_w$. One easily sees that $N(\mathbb{F}(\!(u_0)\!))=W\ltimes T(\mathbb{F}(\!(u_0)\!))$. The Kottwitz homomorphism $\kappa_T$ induces an isomorphism $T(\mathbb{F}(\!(u_0)\!))/T(\mathbb{F}(\!(u_0)\!))_1\xrightarrow{\sim} X$ and we thereby identify $\widetilde{W}$ with $W\ltimes X$. To avoid any confusion of the product inside $\widetilde{W}$ and the canonical action of $S_n$ on $\ZZ^n$, we will always denote the element of $\widetilde{W}$ corresponding to $\lambda\in X$ by $u^\lambda$. Recall from \cite[\textsection 2.5]{gy1} the notion of an extended alcove $(x_i)_{i=0}^{n-1}$ for $\mathrm{GL}_{n}$. An \emph{extended alcove for $\mathrm{GU}$} is an extended alcove $(x_i)_{i=0}^{n-1}$ for $\mathrm{GL}_{n}$ such that \begin{equation*} \exists r\in\ZZ\forall i\in\{0,\dots,n\}\forall j\in\{1,\dots,n\}:\ x_i(j)+x_{n-i}(n+1-j)=2r-1. \end{equation*} Here $x_n=x_0+(1^{(n)})$. Also recall the standard alcove $(\omega_i)_{i=0}^{n-1}$. As in the linear case treated in loc.\ cit., we identify $\widetilde{W}$ with the set of extended alcoves for $\mathrm{GU}$ by using the standard alcove as a base point. Write $\mathbf{e}=(e^{(n)})$. \begin{defn}[{Cf.\ \cite{kottwitz_rapoport}}] An extended alcove $(x_i)_{i=0}^{n-1}$ for $\mathrm{GU}$ is called \emph{permissible} if it satisfies the following conditions for all $i\in\{0,\dots,n-1\}$. \begin{enumerate} \item $\omega_i\leq x_i \leq \omega_i+\mathbf{e}$, where $\leq$ is to be understood componentwise. \item $\sum_{j=1}^{n} x_i(j)=ne_0-i$. \end{enumerate} Denote by $\mathrm{Perm}$ the set of all permissible extended alcoves for $\mathrm{GU}$. \end{defn} \begin{prop} The inclusion $N(\mathbb{F}(\!(u_0)\!))\subset \mathrm{GU}(\mathbb{F}(\!(u_0)\!))$ induces a bijection $\widetilde{W}\xrightarrow{\sim} I_{\mathrm{GU}}(\mathbb{F})\backslash \mathcal \mathrm{GU}(\mathbb{F}(\!(u_0)\!))/I_{\mathrm{GU}}(\mathbb{F})$. In other words, \begin{equation*} \mathrm{GU}(\mathbb{F}(\!(u_0)\!))=\coprod_{x\in\widetilde{W}} I_{\mathrm{GU}}(\mathbb{F})xI_{\mathrm{GU}}(\mathbb{F}). \end{equation*} Under this bijection, the subset \begin{equation*} \Aut(\overline{\mathcal L})(\mathbb{F}[u_0]/u_0^{e_0})\backslash M^{e,n}(\mathbb{F})\subset I_{\mathrm{GU}}(\mathbb{F})\backslash \mathrm{GU}(\mathbb{F}(\!(u_0)\!))/I_{\mathrm{GU}}(\mathbb{F}) \end{equation*} of Proposition \ref{PropIndicesUni} corresponds to the subset $\mathrm{Perm}\subset \widetilde{W}$. \end{prop} \begin{proof} The first statement is discussed in \cite[4.4]{smithling_unitary_odd}, \cite[6.4]{smithling_unitary_even}. The second statement follows easily from the explicit description of the image of $\alpha$ in Proposition \ref{PropLocalModelintoFlagUni}. \end{proof} \begin{cor}\label{CorIndexSetUni} Under the identifications of Theorem \ref{ThmIndicesUni}, the set \allowbreak$\prod_{\sigma\in\mathfrak S} \mathrm{Perm}$ constitutes a set of representatives of $\Aut(\mathcal L)(\mathbb{F})\backslash M^{\mathrm{loc}}(\mathbb{F})$. \end{cor} \begin{remark} As explained in Remark \ref{RemVvsFSym}, the normalization of Corollary \ref{CorIndexSetUni} differs from the one of Definition \ref{DefnIndexKR} by the automorphism $\prod_{\sigma\in\mathfrak S} \mathrm{Perm}\to \prod_{\sigma\in\mathfrak S} \mathrm{Perm},\ (x_\sigma)\mapsto (u^ex_\sigma^{-1})$ \end{remark} \subsection{The \texorpdfstring{$p$}{p}-rank on a KR stratum} Recall from Section \ref{SecLocalModelGen} the scheme $\mathcal A/\mathcal O_{E_\mathcal Q}$ associated with our choice of PEL datum, and the KR stratification \begin{equation*} \mathcal A(\mathbb{F})=\coprod_{x\in \Aut(\mathcal L)(\mathbb{F})\backslash M^{\mathrm{loc}}(\mathbb{F})}\mathcal A_x. \end{equation*} We have identified the occurring index set with $\prod_{\sigma\in\mathfrak S}\mathrm{Perm}$ in Corollary \ref{CorIndexSetUni} . We can then state the following result. \begin{thm}\label{ThmPrankUni} Let $x=(x_\sigma)_\sigma\in \prod_{\sigma\in\mathfrak S} \mathrm{Perm}$. Write $x_\sigma=w_\sigma u^{\lambda_\sigma}$ with $w_\sigma\in W,\ \lambda_\sigma\in X$. Then the $p$-rank on $\mathcal A_x$ is constant with value \begin{equation*} g\cdot |\{1\leq i\leq n\mathrel{|} \forall \sigma\in\mathfrak S(w_\sigma(i)=i\wedge \lambda_\sigma(i)=0)\}|. \end{equation*} \end{thm} \begin{proof} The proof is identical to the one of Theorem \ref{ThmPrank}. \end{proof} \section{The inert unitary case}\label{SecUni2} \subsection{The PEL datum}\label{SecPELUni2} We start with the PEL datum defined in Section \ref{SecPELUni}. We assume that $p\mathcal O_{F_0}=(\mathcal P_0)^{e}$ for a single prime $\mathcal P_0$ of $\mathcal O_{F_0}$ and that $\mathcal P_0\mathcal O_F=\mathcal P$ for a single prime $\mathcal P$ of $\mathcal O_F$. Denote by $f_0=[k_{\mathcal P_0}:\mathbb{F}_p]$ and $f=[k_{\mathcal P}:\mathbb{F}_p]$ the corresponding inertia degrees, so that $f=2f_0$. We fix once and for all a uniformizer $\pi$ of $\mathcal O_{F_0}\otimes\ZZ_{(p)}$. Then $\pi$ is also a uniformizer of $\mathcal O_{F}\otimes\ZZ_{(p)}$. Denote by $\mathfrak C=\mathfrak C_{\mathcal O_{F_\mathcal P}\mathrel{|} \ZZ_p}$ the corresponding inverse different. Choose a generator $\delta$ of $\mathfrak C$ satisfying $\delta^\ast=-\delta$. Consequently the form $\delta^{-1}\left(\cdot,\cdot\right)'_{\QQ_p}:V_{\QQ_p}\times V_{\QQ_p}\to F_\mc P$ is $\ast$-hermitian and we assume that it \emph{splits}, i.e.\ that there is a basis $(e_1,\dots,e_n)$ of $V_{\QQ_p}$ over $F_\mc P$ such that $\pairt{e_i}{e_{n+1-j}}_{\QQ_p}'=\delta\delta_{ij}$ for $1\leq i,j\leq n$. Let $0\leq i<n$. We denote by $\Lambda_i$ the $\mathcal O_{F_\mathcal P}$-lattice in $V_{\QQ_p}$ with basis \begin{equation*} \mathfrak E_i=(\pi^{-1}e_1,\ldots,\pi^{-1}e_i,e_{i+1},\ldots,e_{n}). \end{equation*} For $k\in\ZZ$ we further define $\Lambda_{nk+i}=\pi^{-k}\Lambda_i$ and we denote by $\mathfrak E_{nk+i}$ the corresponding basis obtained from $\mathfrak E_i$. Then $\mathcal L=(\Lambda_i)_i$ is a complete chain of $\mathcal O_{F_\mathcal P}$-lattices in $V_{\QQ_p}$. For $i\in\ZZ$, the dual lattice $\Lambda_i^\vee:=\{x\in V_{\QQ_p}\mathrel{|} \pairt{x}{\Lambda_i}_{\QQ_p}\subset\ZZ_p\}$ of $\Lambda_i$ is given by $\Lambda_{-i}$. Consequently the chain $\mathcal L$ is self-dual. Let $i\in\ZZ$. We denote by $\rho_i:\Lambda_{i}\to \Lambda_{i+1}$ the inclusion, by $\vartheta_i:\Lambda_{n+i}\to \Lambda_i$ the isomorphism given by multiplication with $\pi$ and by $\left(\cdot,\cdot\right)_i:\Lambda_i\times \Lambda_{-i}\to \ZZ_p$ the restriction of $\left(\cdot,\cdot\right)_{\QQ_p}$. Then $(\Lambda_{i},\rho_{i},\vartheta_{i},\left(\cdot,\cdot\right)_{i})_i$ is a polarized chain of $\mathcal O_{F_\mc P}$-modules of type $(\mathcal L)$, which, by abuse of notation, we also denote by $\mathcal L=\mathcal L^{\mathrm{inert}}$. Denote by $\left<\cdot,\cdot\right>_i:\Lambda_i\times \Lambda_{-i}\to \mathcal O_{F_\mathcal P}$ the restriction of the $\ast$-hermitian form $\delta^{-1}\left(\cdot,\cdot\right)'_{\QQ_p}$. It is the $\ast$-sesquilinear form described by the matrix $\widetilde{I}_n$ with respect to $\mathfrak E_i$ and $\mathfrak E_{-i}$. Denote by $\Sigma_0$ the set of all embeddings $F_0\hookrightarrow\RR$ and by $\Sigma$ the set of all embeddings $F\hookrightarrow \CC$. Also write $\mathfrak S=\Gal(k_\mathcal P/\mathbb{F}_p)$ and $\mathfrak S_0=\Gal(k_{\mathcal P_0}/\mathbb{F}_p)$. Let $E'$ be the Galois closure of $F$ inside $\CC$ and choose a prime $\mathcal Q'$ of $E'$ over $\mathcal P$. Consider the maps $\gamma:\Sigma\to \mathfrak S$ and $\gamma_0:\Sigma_0\to \mathfrak S_0$ of Lemma \ref{LemResidueMapinNonGaloisCase}. For each $\sigma\in\mathfrak S_0$ we denote by $\tau_{\sigma,1},\tau_{\sigma,2}\in\mathfrak S$ the two elements with $\tau_{\sigma,j}\big|_{k_{\mathcal P_0}}=\sigma$. Let $\sigma\in\Sigma_0$ and $j\in\{1,2\}$. There is a unique $\tau_{\sigma,j}\in\Sigma$ with $\tau_{\sigma,j}\big|_{F_0}=\sigma$ satisfying \begin{equation}\label{EqInterplayOfGammas} \gamma(\tau_{\sigma,j})=\tau_{\gamma_0(\sigma),j}. \end{equation} Exactly as in Section \ref{SecUni}, we define for each $\sigma\in\Sigma_0$ integers $r_\sigma,s_\sigma$ with $r_\sigma+s_\sigma=n$, and using these the element $J\in \End_{B\otimes \RR}(V\otimes \RR)$. Denote by $V_{-i}$ the $(-i)$-eigenspace of $J_\CC$. As before, we construct an $\mathcal O_F\otimes\mathcal O_{E'}$-module $M_0$ which is finite locally free over $\mathcal O_{E'}$, such that $M_0\otimes_{\mathcal O_{E'}}\CC=V_{-i}$ as $\mathcal O_F\otimes\CC$-modules. \subsection{The special fiber of the determinant morphism} Let $\sigma\in\mathfrak S_0$. We define \begin{equation*} \overline{r}_\sigma=\sum_{\sigma'\in \gamma_0^{-1}(\sigma)} r_{\sigma'}\quad \text{and}\quad \overline{s}_\sigma=\sum_{\sigma'\in \gamma_0^{-1}(\sigma)} s_{\sigma'}. \end{equation*} As the fibers of $\gamma_0$ have cardinality $e$, it follows that $\overline{r}_\sigma+\overline{s}_\sigma=ne$. We fix once and for all an embedding $\iota_{\mathcal Q'}:k_{\mathcal Q'}\hookrightarrow \mathbb{F}$. We consider $\mathbb{F}$ as an $\mathcal O_{E'}$-algebra with respect to the composition $\mathcal O_{E'}\xrightarrow{\rho_{\mathcal Q'}}k_{\mathcal Q'}\overset{\iota_{\mathcal Q'}}{\hookrightarrow} \mathbb{F}$. Also $\iota_{\mathcal Q'}$ induces an embedding $\iota_\mathcal P:k_{\mathcal \mathcal P}\hookrightarrow \mathbb{F}$ and thereby an identification of the set of all embeddings $k_{\mathcal P}\hookrightarrow \mathbb{F}$ with $\mathfrak S$. Our choice of uniformizer $\pi$ induces a canonical isomorphism \begin{equation} \label{EqdecompspecialfiberUni2} \mathcal O_{F}\otimes \mathbb{F}=\prod_{\sigma\in \mathfrak S_0} \mathbb{F}[u]/(u^{e})\times \mathbb{F}[u]/(u^{e}). \end{equation} Here in the component $\mathbb{F}[u]/(u^{e})\times \mathbb{F}[u]/(u^{e})$ corresponding to $\sigma\in \mathfrak S_0$, the first factor is supposed to correspond to $\tau_{\sigma,1}$ and the second factor is supposed to correspond to $\tau_{\sigma,2}$. \begin{prop}\label{PropDeterminantSpecialFiberUni2} Let $x\in \mathcal O_F$ and let $\bigl((q_{\tau_{\sigma,1}},q_{\tau_{\sigma,2}})\bigr)_\sigma\in \prod_{\sigma\in \mathfrak S_0} \mathbb{F}[u]/(u^{e})\times \mathbb{F}[u]/(u^{e})$ be the element corresponding to $x\otimes 1$ under \eqref{EqdecompspecialfiberUni2}. Then \begin{equation*} \chi_\mathbb{F}(x| M_0\otimes_{\mathcal O_{E'}}\mathbb{F})=\prod_{\sigma\in \mathfrak S_0}\bigl(T-q_{\tau_{\sigma,1}}(0)\bigr)^{\overline{s}_\sigma}\bigl(T-q_{\tau_{\sigma,2}}(0)\bigr)^{\overline{r}_\sigma} \end{equation*} in $\mathbb{F}[T]$. \end{prop} \begin{proof} Reduce $\chi_{\mathcal O_{E'}}(x|M_0)$ modulo $\mathcal Q'$, using \eqref{EqInterplayOfGammas}. \end{proof} Denote by $E=\QQ(\tr_\CC(x\otimes 1|V_{-i});\ x\in F)$ the reflex field and define $\mathcal Q=\mathcal Q'\cap \mathcal O_E$. The morphism $\det_{V_{-i}}$ is defined over $\mathcal O_E$. \subsection{The local model} For the chosen PEL datum, Definition \ref{DefnLocalModelGen} amounts to the following. \begin{defn}\label{DefnLocalModelUni2} The local model $M^{\mathrm{loc}}=M^{\mathrm{loc,inert}}$ is the functor on the category of $\mathcal O_{E_\mathcal Q}$-algebras with $M^{\mathrm{loc}}(R)$ the set of tuples $(t_i)_{i\in \ZZ}$ of $\mathcal O_{F}\otimes R$-submodules $t_i\subset \Lambda_{i,R}$, satisfying the following conditions for all $i\in\ZZ$. \renewcommand\theenumi{\alph{enumi}} \begin{enumerate} \item\label{DefnLocalModelUni2-Functoriality} $\rho_{i,R}(t_i)\subset t_{i+1}$. \item\label{DefnLocalModelUni2-Projectivity} The quotient $\Lambda_{i,R}/t_i$ is a finite locally free $R$-module. \item\label{DefnLocalModelUni2-Determinant} We have an equality \begin{equation*} \det_{\Lambda_{i,R}/t_i}=\det_{V_{-i}}\otimes_{\mathcal O_E}R \end{equation*} of morphisms $V_{\mathcal O_F\otimes R}\to \mathbb{A}^1_R$. \item\label{DefnLocalModelUni2-DualityCondition} Under the pairing $\left(\cdot,\cdot\right)_{i,R}:\Lambda_{i,R}\times \Lambda_{-i,R}\to R$, the submodules $t_i$ and $t_{-i}$ pair to zero. \item\label{DefnLocalModelUni2-Periodicity} $\vartheta_i(t_{n+i})=t_i$. \end{enumerate} \end{defn} \subsection{The special fiber of the local model}\label{SecSpecialFibLocModUni2} For $i\in\ZZ$, denote by $\overline{\Lambda}_{i}$ the free $\mathbb{F}[u]/u^{e}$-module $(\mathbb{F}[u]/u^{e})^{n}$ and by $\overline{\mathfrak E}_{i}$ its canonical basis. Denote by $\overline{\vartheta}_{i}:\overline{\Lambda}_{n+i}\to \overline{\Lambda}_{i}$ the identity morphism. Consider the map $\overline{\ast}:\mathbb{F}[u]/u^{e}\times \mathbb{F}[u]/u^{e}\to \mathbb{F}[u]/u^{e}\times \mathbb{F}[u]/u^{e},\ (a,b)\mapsto (b,a)$. Let $\overline{\Lambda}_{i,1}$ and $\overline{\Lambda}_{i,2}$ be two copies of $\overline{\Lambda}_i$ and denote by $\overline{\left<\cdot,\cdot\right>}_{i,1}:\overline{\Lambda}_{i,1}\times \overline{\Lambda}_{-i,2}\to \mathbb{F}[u]/u^{e}$ (resp.\ $\overline{\left<\cdot,\cdot\right>}_{i,2}:\overline{\Lambda}_{i,2}\times \overline{\Lambda}_{-i,1}\to \mathbb{F}[u]/u^{e}$) the perfect bilinear map described by the matrix $\widetilde{I}_n$ with respect to $\overline{\mathfrak E}_{i,1}$ and $\overline{\mathfrak E}_{-i,2}$ (resp.\ $\overline{\mathfrak E}_{i,2}$ and $\overline{\mathfrak E}_{-i,1}$). Consider the pairing \begin{align*} \overline{\left<\cdot,\cdot\right>}_{i}:(\overline{\Lambda}_{i,1}\times \overline{\Lambda}_{i,2})\times (\overline{\Lambda}_{-i,1}\times \overline{\Lambda}_{-i,2})&\to \mathbb{F}[u]/u^{e}\times \mathbb{F}[u]/u^{e},\\ \bigl((x_1,x_2),(y_1,y_2)\bigr)&\mapsto \bigl(\pair{x_1}{y_2}_{i,1},\pair{x_2}{y_1}_{i,2}\bigr). \end{align*} It is a perfect $\overline{\ast}$-sesquilinear pairing. For $k\in\ZZ$ and $0\leq i<n$, let $\overline{\rho}_{nk+i}:\overline{\Lambda}_{nk+i}\to \overline{\Lambda}_{nk+i+1}$ be the morphism described by the matrix $\mathrm{diag}(1^{(i)},u,1^{(n-i-1)})$ with respect to $\overline{\mathfrak E}_{nk+i}$ and $\overline{\mathfrak E}_{nk+i+1}$. \begin{defn}\label{DefnSpecialLocalModelUni2} Let $r,s\in\NN$ with $r+s=ne$. Define a functor $M^{e,n,r}$ on the category of $\mathbb{F}$-algebras with $M^{e,n,r}(R)$ the set of tuples $(t_i)_{i\in \ZZ}$ of $R[u]/u^{e}$-submodules $t_i\subset \overline{\Lambda}_{i,R}$ satisfying the following conditions for all $i\in\ZZ$. \renewcommand\theenumi{\alph{enumi}} \begin{enumerate} \item $\overline{\rho}_{i,R}(t_i)\subset t_{i+1}$. \item The quotient $\overline{\Lambda}_{i,R}/t_i$ is a finite locally free $R$-module. \item\label{DefnSpecialLocalModelUni2-Determinant} For all $p\in R[u]/u^e$, we have \begin{equation*} \chi_R(p|\overline{\Lambda}_{i,R}/t_i)=\bigl(T-p(0)\bigr)^{s} \end{equation*} in $R[T]$. \item $\overline{\vartheta}_i(t_{n+i})=t_i$. \end{enumerate} \end{defn} Let $i\in\ZZ$. From \eqref{EqdecompspecialfiberUni2} we obtain an isomorphism \begin{equation}\label{LambdaInSpecialFiberUni2} \Lambda_{i,\mathbb{F}}=\prod_{\sigma\in \mathfrak S_0}\overline{\Lambda}_{i,1}\times \overline{\Lambda}_{i,2} \end{equation} by identifying the basis $\mathfrak E_{i,\mathbb{F}}$ with the product of the bases $\overline{\mathfrak E}_i$. Under this identification, the morphism $\rho_{i,\mathbb{F}}$ decomposes into the morphisms $\overline{\rho}_i$, the pairing $\left<\cdot,\cdot\right>_{i,\mathbb{F}}$ decomposes into the pairings $\overline{\left<\cdot,\cdot\right>}_i$ and the morphism $\vartheta_{i,\mathbb{F}}$ decomposes into the morphisms $\overline{\vartheta}_i$. Let $R$ be an $\mathbb{F}$-algebra and let $(t_i)_{i\in\ZZ}$ be a tuple of $\mathcal O_F\otimes R$-submodules $t_i\subset \Lambda_{i,R}$. Then \eqref{LambdaInSpecialFiberUni2} induces decompositions $t_i=\prod_{\sigma\in \mathfrak S_0} t_{i,\tau_{\sigma,1}}\times t_{i,\tau_{\sigma,2}}$ into $R[u]/u^{e}$-submodules $t_{i,\tau_{\sigma,j}}\subset\overline{\Lambda}_{i,j,R}$. The following statement is then clear (cf.\ the proof of Proposition \ref{PropDecompositionofLocalModel}). \begin{prop}\label{PropDecompositionofLocalModelUni2} The morphism $\Phi_1:M^{\mathrm{loc}}_\mathbb{F}\to \prod_{\sigma\in \mathfrak S_0} M^{e,n,\overline{r}_\sigma}$ given on $R$-valued points by \begin{align*} M^{\mathrm{loc}}_\mathbb{F}(R)&\to \prod_{\sigma\in \mathfrak S_0} M^{e,n,\overline{r}_\sigma}(R),\\ (t_i)&\mapsto \left((t_{i,\tau_{\sigma,1}})_i\right)_\sigma \end{align*} is an isomorphism of functors on the category of $\mathbb{F}$-algebras. \end{prop} \begin{remark}\label{RemarkDifferentDecompofLocalModel} For symmetry reasons, also the morphism $\Phi_2:M^{\mathrm{loc}}_\mathbb{F}\to \prod_{\sigma\in \mathfrak S_0} M^{e,n,\overline{s}_\sigma}$ given on $R$-valued points by \begin{align*} M^{\mathrm{loc}}_\mathbb{F}(R)&\to \prod_{\sigma\in \mathfrak S_0} M^{e,n,\overline{s}_\sigma}(R),\\ (t_i)&\mapsto \left((t_{i,\tau_{\sigma,2}})_i\right)_\sigma \end{align*} is an isomorphism of functors on the category of $\mathbb{F}$-algebras. The morphism $\prod_{\sigma\in \mathfrak S_0} M^{e,n,\overline{r}_\sigma}\to \prod_{\sigma\in \mathfrak S_0} M^{e,n,\overline{s}_\sigma}$ making commutative the diagram \begin{equation*} \xymatrix{ &\prod_{\sigma\in \mathfrak S_0} M^{e,n,\overline{r}_\sigma}\ar[dd]\\ M^{\mathrm{loc}}_\mathbb{F}\ar[ur]^{\Phi_1}\ar[dr]_{\Phi_2}& \\ &\prod_{\sigma\in \mathfrak S_0} M^{e,n,\overline{s}_\sigma}\\ } \end{equation*} is given by on $R$-valued points by \begin{equation} \begin{aligned} \prod_{\sigma\in \mathfrak S_0} M^{e,n,\overline{r}_\sigma}(R)&\to \prod_{\sigma\in \mathfrak S_0} M^{e,n,\overline{s}_\sigma}(R),\\ ((t_{i,\sigma})_i)_{\sigma}&\mapsto ((t_{-i,\sigma}^{\perp,\overline{\left<\cdot,\cdot\right>}_{-i,1,R}})_i)_{\sigma}. \end{aligned} \label{EqRelationBetweenDifferentMs} \end{equation} \end{remark} \subsection{Embedding the local model into the affine flag variety} Recall from Section \ref{SecLattices} the affine flag variety $\mathcal F$. Let $R$ be an $\mathbb{F}$-algebra. We consider an $R[u]/u^e$-module as an $R[\![u]\!]$-module via the canonical projection $R[\![u]\!]\to R[u]/u^e$. For $i\in\ZZ$, denote by $\alpha_i:\widetilde{\Lambda}_i\to \overline{\Lambda}_{i,R}$ the morphism described by the identity matrix with respect to $\widetilde{\mathfrak E}_i$ and $\overline{\mathfrak E}_i$. It induces an isomorphism $\widetilde{\Lambda}_i/u^e\widetilde{\Lambda}_i\xrightarrow{\sim}\overline{\Lambda}_{i,R}$. Clearly the following diagrams commute. \begin{equation*} \xymatrix{ \widetilde{\Lambda}_i\ar[d]_{\alpha_i}\ar@{}[r]|{\subset}&\widetilde{\Lambda}_{i+1}\ar[d]^{\alpha_{i+1}}\\ \overline{\Lambda}_{i,R}\ar[r]^-{\overline{\rho}_{i,R}}&\overline{\Lambda}_{i+1,R}, } \quad \xymatrix{ \widetilde{\Lambda}_i\ar[d]_{\alpha_i}&\widetilde{\Lambda}_{n+i}\ar[d]^{\alpha_{n+i}}\ar[l]_{u\cdot}\\ \overline{\Lambda}_{i,R}&\overline{\Lambda}_{n+i,R}\ar[l]_{\overline{\vartheta}_{i,R}}. } \end{equation*} Let $r,s\in\NN$ with $r+s=ne$. The following proposition allows us to consider $M^{e,n,r}$ as a subfunctor of $\mathcal F$. \begin{prop}[{\cite[\textsection 4]{pr2}}]\label{PropLocalModelintoFlagUni2} There is an embedding $\alpha:M^{e,n,r}\hookrightarrow \mathcal F$ given on $R$-valued points by \begin{align*} M^{e,n,r}(R)&\to \mathcal F(R),\\ (t_i)_i&\mapsto (\alpha_i^{-1}(t_i))_i. \end{align*} It induces a bijection from $M^{e,n,r}(R)$ onto the set of those $(L_i)_i\in \mathcal F(R)$ satisfying the following conditions for all $i\in\ZZ$. \begin{enumerate} \item $u^e\widetilde{\Lambda}_i\subset L_i\subset\widetilde{\Lambda}_i$. \item For all $p\in R[u]/u^e$, we have \begin{equation*} \chi_R(p|\widetilde{\Lambda}_{i}/L_i)=\bigl(T-p(0)\bigr)^{s} \end{equation*} in $R[T]$. Here $\widetilde{\Lambda}_i/L_i$ is considered as an $R[u]/u^e$-module using (1). \end{enumerate} \end{prop} \begin{proof} Analogous to the proof of Proposition \ref{PropLocalModelintoFlag}. \end{proof} Let $R$ be an $\mathbb{F}$-algebra. Denote by $\widetilde{\left<\cdot,\cdot\right>}:R(\!(u)\!)^n\times R(\!(u)\!)^n\to R(\!(u)\!)$ the bilinear form described by the matrix $\widetilde{I}_n$ with respect to the standard basis of $R(\!(u)\!)^n$ over $R(\!(u)\!)$. Further denote by $\widetilde{\left<\cdot,\cdot\right>}_i:\widetilde{\Lambda}_i\times \widetilde{\Lambda}_{-i}\to R[\![u]\!]$ the restriction of $\widetilde{\left<\cdot,\cdot\right>}$. Note that the diagram \begin{equation*} \xymatrix{ \widetilde{\Lambda}_{i}\times \widetilde{\Lambda}_{-i}\ar[rr]^{\widetilde{\left<\cdot,\cdot\right>}_{i}}\ar[d]_{\alpha_{i}\times \alpha_{-i}}&& R[\![u]\!]\ar[d]\\ \overline{\Lambda}_{i,1,R}\times \overline{\Lambda}_{-i,2,R}\ar[rr]^-{\overline{\left<\cdot,\cdot\right>}_{i,1,R}}&& R[u]/u^e } \end{equation*} commutes. For a lattice $\Lambda$ in $R(\!(u)\!)^{n}$ we define $\Lambda^\vee:=\{x\in R(\!(u)\!)^{n}\mathrel{|} \widetilde{\pair{x}{\Lambda}}\subset R[\![u]\!]\}$. As in Remark \ref{RemarkTwoEmbeddingsUni2}, the morphism $\Psi:M^{e,n,r}\to M^{e,n,s}$ given on $R$-valued points by \begin{align*} M^{e,n,r}(R)&\to M^{e,n,s}(R),\\ (t_i)_i&\mapsto (t_{-i}^{\perp,\overline{\left<\cdot,\cdot\right>}_{-i,1,R}})_i \end{align*} is an isomorphism. \begin{prop}\label{PropDifferentEmbeddingsIntoFlag} The following diagram commutes. \begin{equation*} \xymatrix{ M^{e,n,r}\,\ar@{^{(}->}[r]^-\alpha\ar[d]_-\Psi& \mathcal F\ar[d]^{(L_i)_i \mapsto (u^eL_{-i}^\vee)_i}\\ M^{e,n,s}\,\ar@{^{(}->}[r]^-\alpha& \mathcal F. } \end{equation*} \end{prop} \begin{proof} Similar to the proof of the duality statement in the proof of Proposition \ref{PropLocalModelintoFlag}. \end{proof} Note that $\overline{\mathcal L}=(\overline{\Lambda}_{i},\overline{\rho}_{i},\overline{\vartheta}_{i})$ is a chain of $\mathbb{F}[u]/u^e$-modules of type $(\widetilde{\mathcal L})$. In fact $\overline{\mathcal L}=\widetilde{\mathcal L}\otimes_{\mathbb{F}[\![u]\!]}\mathbb{F}[u]/u^e$. Let $R$ be an $\mathbb{F}$-algebra. There is an obvious action of $\Aut(\overline{\mathcal L})(R[u]/u^e)$ on $M^{e,n,r}(R)$, given by $(\varphi_i)\cdot (t_{i})=(\varphi_i(t_i))$. The canonical morphism $R[\![u]\!]\to R[u]/u^e$ induces a map $\Aut(\widetilde{\mathcal L})(R[\![u]\!])\to \Aut(\overline{\mathcal L})(R[u]/u^e)$ and we thereby extend this $\Aut(\overline{\mathcal L})(R[u]/u^e)$-action on $M^{e,n,r}(R)$ to an $\Aut(\widetilde{\mathcal L})(R[\![u]\!])$-action. \begin{lem}\label{LemDifferentAutsSameOrbitsUni2} Let $R$ be an $\mathbb{F}$-algebra and let $t\in M^{e,n,r}(R)$. We have $\Aut(\widetilde{\mathcal L})(R[\![u]\!])\cdot t=\Aut(\overline{\mathcal L})(R[u]/u^e)\cdot t$. \end{lem} \begin{proof} The map $\Aut(\widetilde{\mathcal L})(R[\![u]\!])\to \Aut(\overline{\mathcal L})(R[u]/u^e)$ is surjective by Proposition \ref{PropLiftingIsosLin}. \end{proof} \begin{lem}\label{LemIvsIprimeUni2} Let $g\in I(\mathbb{F})$. Then $g$ restricts to an automorphism $g_i:\widetilde{\Lambda}_i\xrightarrow{\sim}\widetilde{\Lambda}_i$ for each $i\in\ZZ$. The assignment $g\mapsto (g_i)_i$ defines an isomorphism $I(\mathbb{F})\xrightarrow{\sim} \Aut(\widetilde{\mathcal L})(\mathbb{F}[\![u]\!])$. \end{lem} \begin{proof} Clear (cf.\ the proof of Lemma \ref{LemIvsIprime}). \end{proof} \begin{prop}\label{PropIndicesUni2_2} Let $t\in M^{e,n,r}(\mathbb{F})$. Then $\alpha$ induces a bijection \begin{equation*} \Aut(\overline{\mathcal L})(\mathbb{F}[u]/u^e)\cdot t\xrightarrow{\sim} I(\mathbb{F})\cdot \alpha(t). \end{equation*} Consequently we obtain an embedding \begin{equation*} \Aut(\overline{\mathcal L})(\mathbb{F}[u]/u^e)\backslash M^{e,n,r}(\mathbb{F})\hookrightarrow I(\mathbb{F})\backslash \mathcal F(\mathbb{F}). \end{equation*} \end{prop} \begin{proof} Analogous to the proof of Proposition \ref{PropIndicesSym}. \end{proof} Consider $\alpha':M^{e,n,r}(\mathbb{F})\hookrightarrow \mathcal F(\mathbb{F})\xrightarrow{\phi(\mathbb{F})^{-1}}\Lf \mathrm{GL}_n(\mathbb{F})/I(\mathbb{F})$. \begin{prop}\label{PropIndicesUni2} Let $t\in M^{e,n,r}(\mathbb{F})$. Then $\alpha'$ induces a bijection \begin{equation*} \Aut(\overline{\mathcal L})(\mathbb{F}[u]/u^e)\cdot t\xrightarrow{\sim} I(\mathbb{F})\cdot \alpha'(t). \end{equation*} Consequently we obtain an embedding \begin{equation}\label{Eqalphabar} \Aut(\overline{\mathcal L})(\mathbb{F}[u]/u^e)\backslash M^{e,n,r}(\mathbb{F})\hookrightarrow I(\mathbb{F})\backslash \mathrm{GL}_n(\mathbb{F}(\!(u)\!))/I(\mathbb{F}). \end{equation} \end{prop} \begin{proof} Clear from Proposition \ref{PropIndicesUni2_2}, as the isomorphism $\phi(\mathbb{F})$ is in particular $I(\mathbb{F})$-equivariant. \end{proof} Denote by $\tau$ the adjoint involution for $\widetilde{\left<\cdot,\cdot\right>}$ on $\mathrm{GL}_n(\mathbb{F}(\!(u)\!))$, so that for $g\in\mathrm{GL}_n(\mathbb{F}(\!(u)\!))$ we have $\widetilde{\pair{gx}{y}}=\widetilde{\pair{x}{g^\tau}},\ x,y\in\mathbb{F}(\!(u)\!)^n$. \begin{prop}\label{PropDifferentEmbeddingsIntoFlag2} The vertical maps in the following diagram are well-defined bijections and the diagram commutes. \begin{equation*} \xymatrix{ \Aut(\overline{\mathcal L})(\mathbb{F}[u]/u^e)\backslash M^{e,n,r}(\mathbb{F})\,\ar@{^{(}->}[r]^-{\eqref{Eqalphabar}}\ar[d]_\Psi& I(\mathbb{F})\backslash \mathrm{GL}_n(\mathbb{F}(\!(u)\!))/I(\mathbb{F})\ar[d]^{g\mapsto u^e(g^\tau)^{-1}}\\ \Aut(\overline{\mathcal L})(\mathbb{F}[u]/u^e)\backslash M^{e,n,s}(\mathbb{F})\,\ar@{^{(}->}[r]^-{\eqref{Eqalphabar}}& I(\mathbb{F})\backslash \mathrm{GL}_n(\mathbb{F}(\!(u)\!))/I(\mathbb{F}). } \end{equation*} \end{prop} \begin{proof} In view of Proposition \ref{PropDifferentEmbeddingsIntoFlag} it suffices to note the following statement, which follows from a short computation: Let $\Lambda$ be a lattice in $\mathbb{F}(\!(u)\!)^n$ and let $g\in \mathrm{GL}_n(\mathbb{F}(\!(u)\!))$. Then $(g\Lambda)^\vee=(g^\tau)^{-1}(\Lambda^\vee)$. \end{proof} Let $R$ be an $\mathbb{F}$-algebra and $\varphi=(\varphi_i)_i\in \Aut(\mathcal L)(R)$. The decomposition \eqref{LambdaInSpecialFiberUni2} induces for each $i$ a decomposition of $\varphi_i:\Lambda_{i,R}\xrightarrow{\sim}\Lambda_{i,R}$ into the product of $R[u]/u^e$-linear automorphisms $\varphi_{i,\tau_{\sigma,j}}:\overline{\Lambda}_{i,j,R}\xrightarrow{\sim}\overline{\Lambda}_{i,j,R}$. The following statement is then clear (cf.\ the proof of Proposition \ref{PropDecompositionofLocalModel}). \begin{prop}\label{PropChainMorphismDecomposesUni2} Let $R$ be an $\mathbb{F}$-algebra. The following map is an isomorphism, functorial in $R$. \begin{align*} \Aut(\mathcal L)(R)&\to \prod_{\sigma \in\mathfrak S_0} \Aut(\overline{\mathcal L})(R[u]/u^e),\\ (\varphi_i)_i&\mapsto ((\varphi_{i,\tau_{\sigma,1}})_i)_\sigma. \end{align*} \end{prop} Consider the composition \begin{equation*} \tilde{\alpha}_1:M^{\mathrm{loc}}(\mathbb{F})\xrightarrow{\Phi_1} \prod_{\sigma\in\mathfrak S_0} M^{e,n,\overline{r}_\sigma}(\mathbb{F})\xrightarrow{\prod_\sigma \alpha'} \prod_{\sigma\in\mathfrak S_0} \Lf \mathrm{GL}_n(\mathbb{F})/I(\mathbb{F}). \end{equation*} For $\sigma\in \mathfrak S_0$ denote by $\tilde{\alpha}_{1,\sigma}:M^{\mathrm{loc}}(\mathbb{F})\to \Lf \mathrm{GL}_n(\mathbb{F})/I(\mathbb{F})$ the corresponding component of $\tilde{\alpha}_1$. \begin{thm}\label{ThmIndicesUni2} Let $t\in M^{\mathrm{loc}}(\mathbb{F})$. Then $\tilde{\alpha}_1$ induces a bijection \begin{equation*} \Aut(\mathcal L)(\mathbb{F})\cdot t\xrightarrow{\sim} \prod_{\sigma\in\mathfrak S_0} I(\mathbb{F})\cdot \tilde{\alpha}_{1,\sigma}(t). \end{equation*} Consequently we obtain an embedding \begin{equation*} \iota_1:\Aut(\mathcal L)(\mathbb{F})\backslash M^{\mathrm{loc}}(\mathbb{F})\hookrightarrow \prod_{\sigma\in\mathfrak S_0} I(\mathbb{F})\backslash \mathcal \mathrm{GL}_n(\mathbb{F}(\!(u)\!))/I(\mathbb{F}). \end{equation*} \end{thm} \begin{proof} Identical to the proof of Theorem \ref{ThmIndicesSym}. \end{proof} \begin{remark}\label{RemarkTwoEmbeddingsUni2} In the same way, the composition \begin{equation*} \tilde{\alpha}_2:M^{\mathrm{loc}}(\mathbb{F})\xrightarrow{\Phi_2} \prod_{\sigma\in\mathfrak S_0} M^{e,n,\overline{s}_\sigma}(\mathbb{F})\xrightarrow{\prod_\sigma \alpha'} \prod_{\sigma\in\mathfrak S_0} \Lf \mathrm{GL}_n(\mathbb{F})/I(\mathbb{F}) \end{equation*} induces an embedding \begin{equation*} \iota_2:\Aut(\mathcal L)(\mathbb{F})\backslash M^{\mathrm{loc}}(\mathbb{F})\hookrightarrow \prod_{\sigma\in\mathfrak S_0} I(\mathbb{F})\backslash \mathcal \mathrm{GL}_n(\mathbb{F}(\!(u)\!))/I(\mathbb{F}). \end{equation*} By Proposition \ref{PropDifferentEmbeddingsIntoFlag2} the following diagram commutes. \begin{equation*} \xymatrix{ & \prod_{\sigma\in\mathfrak S_0} I(\mathbb{F})\backslash \mathcal \mathrm{GL}_n(\mathbb{F}(\!(u)\!))/I(\mathbb{F})\ar[dd]^{(g_\sigma)_\sigma\mapsto (u^e(g_\sigma^\tau)^{-1})_\sigma}\\ \Aut(\mathcal L)(\mathbb{F})\backslash M^{\mathrm{loc}}(\mathbb{F})\ar[ur]^{\iota_1}\ar[dr]_{\iota_2}& \\ & \prod_{\sigma\in\mathfrak S_0} I(\mathbb{F})\backslash \mathcal \mathrm{GL}_n(\mathbb{F}(\!(u)\!))/I(\mathbb{F}) } \end{equation*} \end{remark} \subsection{The extended affine Weyl group}\label{SecExtAffWeylUni2} Let $T$ be the maximal torus of diagonal matrices in $\mathrm{GL}_n$ and let $N$ be its normalizer. We denote by $\widetilde{W}=N(\mathbb{F}(\!(u)\!))/T(\mathbb{F}[\![u]\!])$ the extended affine Weyl group of $\mathrm{GL}_n$ with respect to $T$. Setting $W=S_n$ and $X=\ZZ^n$, the group homomorphism $\upsilon:W\ltimes X\to N(\mathbb{F}(\!(u)\!)),\ (w,\lambda)\mapsto A_wu^\lambda$ induces an isomorphism $W\ltimes X\xrightarrow{\sim} \widetilde{W}$. We use it to identify $\widetilde{W}$ with $W\ltimes X$ and consider $\widetilde{W}$ as a subgroup of $\mathrm{GL}_n(\mathbb{F}(\!(u)\!)\!)$ via $\upsilon$. To avoid any confusion of the product inside $\widetilde{W}$ and the canonical action of $S_n$ on $\ZZ^n$, we will always denote the element of $\widetilde{W}$ corresponding to $\lambda\in X$ by $u^\lambda$. Recall from \cite[\textsection 2.5]{gy1} the notion of an extended alcove $(x_i)_{i=0}^{n-1}$ for $\mathrm{GL}_{n}$. Also recall the standard alcove $(\omega_i)_i$. As in loc.\ cit.\ we identify $\widetilde{W}$ with the set of extended alcoves by using the standard alcove as a base point. Let $r,s\in\NN$ with $r+s=ne$ and write $\mathbf{e}=(e^{(n)})$. \begin{defn}[{Cf.\ \cite{kottwitz_rapoport}, \cite[Definition 2.4]{gy1}}] An extended alcove $(x_i)_{i=0}^{n-1}$ is called \emph{$r$-permissible} if it satisfies the following conditions for all $i\in\{0,\dots,n-1\}$. \begin{enumerate} \item $\omega_i\leq x_i \leq \omega_i+\mathbf{e}$, where $\leq$ is to be understood componentwise. \item $\sum_{j=1}^{n} x_i(j)=s-i$. \end{enumerate} Denote by $\mathrm{Perm}_{r}$ the set of all $r$-permissible extended alcoves. \end{defn} \begin{prop} The inclusion $N(\mathbb{F}(\!(u)\!))\subset \mathrm{GL}_n(\mathbb{F}(\!(u)\!))$ induces a bijection $\widetilde{W}\xrightarrow{\sim} I(\mathbb{F})\backslash \mathcal \mathrm{GL}_n(\mathbb{F}(\!(u)\!))/I(\mathbb{F})$. In other words, \begin{equation*} \mathrm{GL}_n(\mathbb{F}(\!(u)\!))=\coprod_{x\in\widetilde{W}} I(\mathbb{F})xI(\mathbb{F}). \end{equation*} Under this bijection, the subset \begin{equation*} \Aut(\overline{\mathcal L})(\mathbb{F}[u]/u^e)\backslash M^{e,n,r}(\mathbb{F})\subset I(\mathbb{F})\backslash \mathrm{GL}_n(\mathbb{F}(\!(u)\!))/I(\mathbb{F}) \end{equation*} of \eqref{Eqalphabar} corresponds to the subset $\mathrm{Perm}_{r}\subset \widetilde{W}$. \end{prop} \begin{proof} The first statement is the well-known Iwahori decomposition. The second statement follows easily from the explicit description of the image of $\alpha$ in Proposition \ref{PropLocalModelintoFlagUni2}. \end{proof} \begin{cor}\label{CorIndexSetUni2} With respect to the embedding $\iota_1$ of Theorem \ref{ThmIndicesUni2}, the set $\prod_{\sigma\in\mathfrak S_0} \mathrm{Perm}_{\bar{r}_\sigma}$ constitutes a set of representatives of $\Aut(\mathcal L)(\mathbb{F})\backslash M^{\mathrm{loc}}(\mathbb{F})$. \end{cor} The following lemma will be used below. \begin{lem}\label{Lemxprime} Let $x\in\widetilde{W}$. Write $x=wu^\lambda$ with $w\in W,\ \lambda\in X$. Define $w'\in W$ and $\lambda'\in X$ by \begin{equation*} w'(i)=n+1-w(n+1-i),\quad 1\leq i\leq n \end{equation*} and \begin{equation*} \lambda'(i)=e-\lambda(n+1-i),\quad 1\leq i\leq n. \end{equation*} Let $x'=w'u^{\lambda'}$. Then $x'=u^e(x^\tau)^{-1}$. \end{lem} \begin{proof} This is an easy computation. \end{proof} \subsection{The \texorpdfstring{$p$}{p}-rank on a KR stratum} Recall from Section \ref{SecLocalModelGen} the scheme $\mathcal A/\mathcal O_{E_\mathcal Q}$ associated with our choice of PEL datum, and the KR stratification \begin{equation*} \mathcal A(\mathbb{F})=\coprod_{x\in \Aut(\mathcal L)(\mathbb{F})\backslash M^{\mathrm{loc}}(\mathbb{F})}\mathcal A_x. \end{equation*} We have identified the occurring index set with $\prod_{\sigma\in\mathfrak S_0}\mathrm{Perm}_{\overline{r}_\sigma}$ in Corollary \ref{CorIndexSetUni2}. We can then state the following result. \begin{thm}\label{ThmPrankUni2} Let $x=(x_\sigma)_\sigma\in \prod_{\sigma\in\mathfrak S_0}\mathrm{Perm}_{\overline{r}_\sigma}$. Write $x_\sigma=w_\sigma u^{\lambda_\sigma}$ with $w_\sigma\in W,\ \lambda_\sigma\in X$ and define elements $w'_\sigma\in W$ and $\lambda_\sigma'\in X$ as in Lemma \ref{Lemxprime}. Then the $p$-rank on $\mathcal A_x$ is constant with value \begin{equation*} g\cdot \left|\left\{1\leq i\leq n\middle\vert \forall \sigma\in\mathfrak S_0\left( \begin{aligned} w_\sigma(i)=w'_\sigma(i)=i\ \wedge\\ \lambda_\sigma(i)=\lambda'_\sigma(i)=0 \end{aligned}\right)\right\}\right|. \end{equation*} \end{thm} \begin{proof} Follows from Proposition \ref{PropDifferentEmbeddingsIntoFlag2} and Lemma \ref{Lemxprime} by the arguments of the proof of Theorem \ref{ThmPrank}. \end{proof} \section{The split unitary case}\label{SecUni3} \subsection{The PEL datum}\label{SecPELUni3} We start with the PEL datum defined in Section \ref{SecPELUni}. We assume that $p\mathcal O_{F_0}=(\mathcal P_0)^{e}$ for a single prime $\mathcal P_0$ of $\mathcal O_{F_0}$ and that $\mathcal P_0\mathcal O_F=\mathcal P_+\mathcal P_-$ for two distinct primes $\mathcal P_\pm$ of $\mathcal O_F$. Consequently $\mathcal P_-=(\mathcal P_+)^\ast$. Denote by $f_0=[k_{\mathcal P_0}:\mathbb{F}_p]$ the corresponding inertia degree. We fix once and for all a uniformizer $\pi_0$ of $\mathcal O_{F_0}\otimes\ZZ_{(p)}$. For typographical reasons, we denote the ring of integers in $(F_0)_{\mathcal P_0}$ by $\mathcal O_{\mathcal P_0}$. The inclusion $\mathcal O_{F_0}\hookrightarrow \mathcal O_F$ induces identifications \begin{equation} \label{EqSplittingOfCompletion} \begin{aligned} \mathcal O_{F}\otimes\ZZ_p&= \mathcal O_{\mathcal P_0}\times \mathcal O_{\mathcal P_0},\\ F\otimes\QQ_p&= (F_0)_{\mathcal P_0}\times (F_0)_{\mathcal P_0}. \end{aligned} \end{equation} Here the first (resp.\ second) factor is always supposed to correspond to $\mathcal P_+$ (resp.\ $\mathcal P_-$). Under \eqref{EqSplittingOfCompletion}, the base-change $F\otimes\QQ_p\to F\otimes\QQ_p$ of $\ast$ takes the simple form $(F_0)_{\mathcal P_0}\times (F_0)_{\mathcal P_0}\to (F_0)_{\mathcal P_0}\times (F_0)_{\mathcal P_0},\ (a,b)\mapsto (b,a)$. The identification \eqref{EqSplittingOfCompletion} further induces a decomposition $V\otimes\QQ_p=V_+\times V_-$ into $(F_0)_{\mathcal P_0}$-vector spaces $V_\pm$. The pairing $\left(\cdot,\cdot\right)'_{\QQ_p}$ decomposes into its restrictions $\left(\cdot,\cdot\right)_\pm:V_\pm\times V_\mp\to (F_0)_{\mathcal P_0}$. Both $\left(\cdot,\cdot\right)_+$ and $\left(\cdot,\cdot\right)_-$ are perfect $(F_0)_{\mathcal P_0}$-bilinear pairings and they are related by the equation $\pairt{v}{w}_+=-\pairt{w}{v}_-,\ v\in V_+, w\in V_-$. Denote by $\mathfrak C_0=\mathfrak C_{\mathcal O_{\mathcal P_0}\mathrel{|} \ZZ_p}$ the corresponding inverse different and fix a generator $\delta_0$ of $\mathfrak C_0$. We fix bases $(e_{1,\pm},\dots,e_{n,\pm})$ of $V_\pm$ over $(F_0)_{\mc P_0}$ such that $\pairt{e_{i,+}}{e_{n+1-j,-}}_+=\delta_0 \delta_{ij}$ for $1\leq i,j\leq n$. Let $0\leq i< n$. We denote by $\Lambda_{i,\pm}$ the $\mathcal O_{\mathcal P_0}$-lattice in $V_{\pm}$ with basis \begin{gather} \mathfrak E_{i,\pm}=(\pi_0^{-1}e_{1,\pm},\ldots,\pi_0^{-1}e_{i,\pm},e_{i+1,\pm},\ldots,e_{n,\pm}). \end{gather} For $k\in\ZZ$ we further define $\Lambda_{nk+i,\pm}=\pi_0^{-k}\Lambda_{i,\pm}$ and we denote by $\mathfrak E_{nk+i,\pm}$ the corresponding basis obtained from $\mathfrak E_{i,\pm}$. Then $\mathcal L_\pm=(\Lambda_{i,\pm})_i$ is a complete chain of $\mathcal O_{\mathcal P_0}$-lattices in $V_\pm$. Let $i\in\ZZ$. We denote by $\rho_{i,\pm}:\Lambda_{i,\pm}\to \Lambda_{i+1,\pm}$ the inclusion and by $\vartheta_{i,\pm}:\Lambda_{n+i,\pm}\to \Lambda_{i,\pm}$ the isomorphism given by multiplication with $\pi_0$. Then $(\Lambda_{i,\pm},\rho_{i,\pm},\vartheta_{i,\pm})$ is a chain of $\mathcal O_{\mathcal P_0}$-modules of type $(\mathcal L_\pm)$ which, by abuse of notation, we also denote by $\mathcal L_\pm$. For $(i,j)\in\ZZ\times\ZZ$ we define $\Lambda_{(i,j)}:=\Lambda_{i,+}\times \Lambda_{j,-}$. Then $\Lambda_{(i,j)}$ is an $\mathcal O_F\otimes\ZZ_p$-lattice in $V_{\QQ_p}$. A basis $\mathfrak E_{(i,j)}$ of $\Lambda_{(i,j)}$ over $\mathcal O_F\otimes\ZZ_p$ is given by the diagonal in $\mathfrak E_{i,+}\times \mathfrak E_{j,-}$. Then $\mathcal L=(\Lambda_{i,j})_{(i,j)}$ is a complete multichain of $\mathcal O_{F}\otimes \ZZ_p$-lattices in $V_{\QQ_p}$. For $(i,j)\in\ZZ\times\ZZ$ the dual lattice $\Lambda_{(i,j)}^\vee:=\{x\in V_{\QQ_p}\mathrel{|} \pairt{x}{\Lambda_{(i,j)}}_{\QQ_p}\subset\ZZ_p\}$ of $\Lambda_{(i,j)}$ is given by $\Lambda_{(-j,-i)}$. Consequently the multichain $\mathcal L$ is a self-dual. Let $(i,j)\in\ZZ\times\ZZ$. We denote by $\rho_{(i,j),+}:\Lambda_{(i,j)}\to \Lambda_{(i+1,j)},\ \rho_{(i,j),-}:\Lambda_{(i,j)}\to \Lambda_{(i,j+1)}$ and $\rho_{(i,j)}:\Lambda_{(i,j)}\to \Lambda_{(i+1,j+1)}$ the inclusions. We denote by $\vartheta_{(i,j),+}:\Lambda_{(n+i,j)}\to \Lambda_{(i,j)}$ (resp.\ $\vartheta_{(i,j),-}:\Lambda_{(i,n+j)}\to \Lambda_{(i,j)}$, resp.\ $\vartheta_{(i,j)}:\Lambda_{(n+i,n+j)}\to \Lambda_{(i,j)}$) the isomorphism given by multiplication with $\pi_0$ in the first (resp.\ second, resp.\ first and second) component. We further denote by $\left(\cdot,\cdot\right)_{(i,j)}:\Lambda_{(i,j)}\times \Lambda_{(-j,-i)}\to \ZZ_p$ the restriction of $\left(\cdot,\cdot\right)_{\QQ_p}$. We find that $(\mathcal L_+,\mathcal L_-)$, equipped with $(\left(\cdot,\cdot\right)_{(i,j)})_{(i,j)}$, is a polarized multichain of $\mathcal O_F\otimes \ZZ_p$-modules of type $(\mathcal L)$, which, by abuse of notation, we also denote by $\mathcal L=\mathcal L^{\mathrm{split}}$. Denote by $\left<\cdot,\cdot\right>_{(i,j)}:\Lambda_{(i,j)}\times \Lambda_{(-j,-i)}\to \mathcal O_{\mathcal P_0}\times \mathcal O_{\mathcal P_0}$ the restriction of the $\ast$-hermitian form $(\delta_0^{-1},-\delta_0^{-1})\left(\cdot,\cdot\right)'_{\QQ_p}$. It is the $\ast$-sesquilinear form described by the matrix $\widetilde{I}_n$ with respect to $\mathfrak E_{(i,j)}$ and $\mathfrak E_{(-j,-i)}$. Denote by $\Sigma_0$ the set of all embeddings $F_0\hookrightarrow\RR$ and by $\Sigma$ the set of all embedding $F\hookrightarrow \CC$. The inclusion $\mathcal O_{F_0}\hookrightarrow \mathcal O_F$ induces an identification of $k_{\mathcal P_\pm}/\mathbb{F}_p$ with $k_{\mathcal P_0}/\mathbb{F}_p$. We write $\mathfrak S_0=\Gal(k_{\mathcal P_0}/\mathbb{F}_p)$ and also identify $\Gal(k_{\mathcal P_\pm}/\mathbb{F}_p)$ with $\mathfrak S_0$. Let $E'$ be the Galois closure of $F$ inside $\CC$ and choose a prime $\mathcal Q'$ of $E'$ over $\mathcal P_+$. Consider the decomposition $\Sigma=\Sigma_+\amalg \Sigma_-$ and the maps $\gamma_0:\Sigma_0\to \mathfrak S_0,\ \gamma_{\pm}:\Sigma_{\pm}\to \mathfrak S_0$ of Lemma \ref{LemGammaSplit}. For $\sigma\in \Sigma_0$ we denote by $\tau_{\sigma,\pm}$ the unique lift of $\sigma$ to $\Sigma_\pm$. Exactly as in Section \ref{SecUni}, we define for each $\sigma\in\Sigma_0$ integers $r_\sigma,s_\sigma$ with $r_\sigma+s_\sigma=n$,\footnote {In Section \ref{SecUni} we have written $\tau_{\sigma,1}$ and $\tau_{\sigma,2}$ instead of $\tau_{\sigma,+}$ and $\tau_{\sigma,-}$, respectively.} and using these the element $J\in\End_{B\otimes\RR}(V\otimes\RR)$. Denote by $V_{\CC,-i}$ the $(-i)$-eigenspace of $J_\CC$. As before, we construct an $\mathcal O_F\otimes\mathcal O_{E'}$-module $M_0$ which is finite locally free over $\mathcal O_{E'}$, such that $M_0\otimes_{\mathcal O_{E'}}\CC=V_{\CC,-i}$ as $\mathcal O_F\otimes\CC$-modules. \subsection{The special fiber of the determinant morphism} For $\sigma\in\mathfrak S_0$ we write \begin{equation*} \overline{r}_\sigma=\sum_{\sigma'\in \gamma_0^{-1}(\sigma)} r_{\sigma'}\quad\text{and}\quad \overline{s}_\sigma=\sum_{\sigma'\in \gamma_0^{-1}(\sigma)} s_{\sigma'}. \end{equation*} As the fibers of $\gamma_0$ have cardinality $e$, it follows that $\overline{r}_\sigma+\overline{s}_\sigma=ne$. We fix once and for all an embedding $\iota_{\mathcal Q'}:k_{\mathcal Q'}\hookrightarrow \mathbb{F}$. We consider $\mathbb{F}$ as an $\mathcal O_{E'}$-algebra with respect to the composition $\mathcal O_{E'}\xrightarrow{\rho_{\mathcal Q'}}k_{\mathcal Q'}\overset{\iota_{\mathcal Q'}}{\hookrightarrow} \mathbb{F}$. Also $\iota_{\mathcal Q'}$ induces an embedding $\iota_{\mathcal P_0}:k_{\mathcal P_0}\hookrightarrow \mathbb{F}$ and thereby an identification of the set of all embeddings $k_{\mathcal P_0}\hookrightarrow \mathbb{F}$ with $\mathfrak S_0$. Consider the isomorphism \begin{equation} \label{EqdecompspecialfiberUni3} \mathcal O_{F}\otimes \mathbb{F}=\prod_{\sigma\in \mathfrak S_0} \mathbb{F}[u]/(u^{e})\times \mathbb{F}[u]/(u^{e}) \end{equation} obtained from \eqref{EqSplittingOfCompletion} and our choice of uniformizer $\pi_0$. \begin{prop}\label{PropDeterminantSpecialFiberUni3} Let $x\in \mathcal O_F$ and let $\bigl((q_{\sigma,+},q_{\sigma,-})\bigr)_\sigma\in \prod_{\sigma\in \mathfrak S_0} \mathbb{F}[u]/(u^{e})\times \mathbb{F}[u]/(u^{e})$ be the element corresponding to $x\otimes 1$ under \eqref{EqdecompspecialfiberUni3}. Then \begin{equation*} \chi_\mathbb{F}(x| M_0\otimes_{\mathcal O_{E'}}\mathbb{F})=\prod_{\sigma\in \mathfrak S_0}\bigl(T-q_{\sigma,+}(0)\bigr)^{\overline{s}_\sigma}\bigl(T-q_{\sigma,-}(0)\bigr)^{\overline{r}_\sigma} \end{equation*} in $\mathbb{F}[T]$. \end{prop} \begin{proof} Reduce $\chi_{\mathcal O_{E'}}(x|M_0)$ modulo $\mathcal Q'$, using \eqref{EqRedOfSigmaUnify}. \end{proof} Denote by $E=\QQ(\tr_\CC(x\otimes 1|V_{-i});\ x\in F)$ the reflex field and define $\mathcal Q=\mathcal Q'\cap \mathcal O_E$. The morphism $\det_{V_{-i}}$ is defined over $\mathcal O_E$. \subsection{The local model} For the chosen PEL datum, Definition \ref{DefnLocalModelGen} amounts to the following. \begin{defn}\label{DefnLocalModelUni3} The local model $M^{\mathrm{loc}}=M^{\mathrm{loc,split}}$ is the functor on the category of $\mathcal O_{E_\mathcal Q}$-algebras with $M^{\mathrm{loc}}(R)$ the set of tuples $(t_{(i,j)})_{(i,j)\in \ZZ\times\ZZ}$ of $\mathcal O_{F}\otimes R$-submodules $t_{(i,j)}\subset \Lambda_{(i,j),R}$ satisfying the following conditions for all $(i,j)\in\ZZ\times \ZZ$. \renewcommand\theenumi{\alph{enumi}} \begin{enumerate} \item\label{DefnLocalModelUni3-Functoriality} $\rho_{(i,j),+,R}(t_{(i,j)})\subset t_{(i+1,j)}$ and $\rho_{(i,j),-,R}(t_{(i,j)})\subset t_{(i,j+1)}$. \item\label{DefnLocalModelUni3-Projectivity} The quotient $\Lambda_{(i,j),R}/t_{(i,j)}$ is a finite locally free $R$-module. \item\label{DefnLocalModelUni3-Determinant} We have an equality \begin{equation*} \det_{\Lambda_{(i,j),R}/t_{(i,j)}}=\det_{V_{-i}}\otimes_{\mathcal O_E}R \end{equation*} of morphisms $V_{\mathcal O_F\otimes R}\to \mathbb{A}^1_R$. \item\label{DefnLocalModelUni3-DualityCondition} Under the pairing $\left(\cdot,\cdot\right)_{(i,j),R}:\Lambda_{(i,j),R}\times \Lambda_{(-j,-i),R}\to R$, the submodules $t_{(i,j)}$ and $t_{(-j,-i)}$ pair to zero. \item\label{DefnLocalModelUni3-Periodicity} $\vartheta_{(i,j),+,R}(t_{(n+i,j)})=t_{(i,j)}$ and $\vartheta_{(i,j),-,R}(t_{(i,n+j)})=t_{(i,j)}$. \end{enumerate} \end{defn} \begin{remark}\label{RemarkLocModDecompUni3} Let $R$ be an $\mathcal O_{E_\mathcal Q}$-algebra and let $(t_{(i,j)})_{(i,j)}\in M^{\mathrm{loc}}(R)$. For $(i,j)\in\ZZ\times\ZZ$, the decomposition \eqref{EqSplittingOfCompletion} induces a decomposition $t_{(i,j)}=t_{(i,j),+}\times t_{(i,j),-}$ into $\mathcal O_{\mathcal P_0}\otimes_{\ZZ_p} R$-submodules $t_{(i,j),+}\subset \Lambda_{i,+,R}$ and $t_{(i,j),-}\subset \Lambda_{j,-,R}$. As in Remark \ref{RemarkDecompLocalModel} one sees that $t_{(i,j),+}$ (resp.\ $t_{(i,j),-}$) is independent of $j$ (resp.\ $i$). Writing $t_{i,+}=t_{(i,j),+}$ and $t_{j,-}=t_{(i,j),-}$, the tuple $(t_{(i,j)})_{(i,j)}$ is determined by the pair of tuples $((t_{i,+})_i, (t_{j,-})_j)$. \end{remark} Recall from Section \ref{SecUni2} the chain $\mathcal L^{\mathrm{inert}}$ and the functor $M^{\mathrm{loc,inert}}$. The identifications \eqref{EqdecompspecialfiberUni2} and \eqref{EqdecompspecialfiberUni3}, together with our choices of bases, give rise to a canonical identification of the tuple $(\Lambda_{(i,i),\mathbb{F}},\rho_{(i,i),\mathbb{F}},\vartheta_{(i,i),\mathbb{F}},\left(\cdot,\cdot\right)_{(i,i),\mathbb{F}})_i$ with the chain $\mathcal L^{\mathrm{inert}}\otimes_{\ZZ_p}\mathbb{F}$. We can then state the following result. \begin{prop}\label{PropMSpinvsMLoc} \begin{enumerate} \item The morphism $M^{\mathrm{loc,split}}_\mathbb{F}\to M^{\mathrm{loc,inert}}_\mathbb{F}$ given on $R$-valued points by \begin{align*} M^{\mathrm{loc,split}}_\mathbb{F}(R)&\to M^{\mathrm{loc,inert}}_\mathbb{F}(R),\\ (t_{(i,j)})_{(i,j)}&\mapsto (t_{(i,i)})_i \end{align*} is an isomorphism. \item The morphism $\Aut(\mathcal L^{\mathrm{split}})_\mathbb{F}\to \Aut(\mathcal L^{\mathrm{inert}})_\mathbb{F}$ given on $R$-valued points by \begin{align*} \Aut(\mathcal L^{\mathrm{split}})_\mathbb{F}(R)&\to \Aut(\mathcal L^{\mathrm{inert}})_\mathbb{F}(R),\\ (\varphi_{(i,j)})_{(i,j)}&\mapsto (\varphi_{(i,i)})_i \end{align*} is an isomorphism. \end{enumerate} \end{prop} \begin{proof} Clear in view of Remark \ref{RemarkLocModDecompUni3} and Propositions \ref{PropDeterminantSpecialFiberUni2}, \ref{PropDeterminantSpecialFiberUni3}. \end{proof} Consequently all the statements about $M^{\mathrm{loc,inert}}_\mathbb{F}$ from Section \ref{SecUni2} are also valid for $M^{\mathrm{loc,split}}_\mathbb{F}$. \subsection{The \texorpdfstring{$p$}{p}-rank on a KR stratum} Recall from Section \ref{SecLocalModelGen} the scheme $\mathcal A/\mathcal O_{E_\mathcal Q}$ associated with our choice of PEL datum, and the KR stratification \begin{equation*} \mathcal A(\mathbb{F})=\coprod_{x\in \Aut(\mathcal L)(\mathbb{F})\backslash M^{\mathrm{loc}}(\mathbb{F})}\mathcal A_x. \end{equation*} We have identified the occurring index set with $\prod_{\sigma\in\mathfrak S_0}\mathrm{Perm}_{\overline{r}_\sigma}$ in Corollary \ref{CorIndexSetUni2}. We can then state the following result. \begin{thm}\label{ThmPrankUni3} Let $x=(x_\sigma)_\sigma\in \prod_{\sigma\in\mathfrak S_0}\mathrm{Perm}_{\overline{r}_\sigma}$. Write $x_\sigma=w_\sigma u^{\lambda_\sigma}$ with $w_\sigma\in W,\ \lambda_\sigma\in X$. Then the $p$-rank on $\mathcal A_{x}$ is constant with value \begin{align*} &\phantom{{}+{}}g_0\cdot |\{1\leq i\leq n\mathrel{|} \forall \sigma\in\mathfrak S_0(w_\sigma(i)=i \wedge \lambda_\sigma(i)=0)\}|\\ &+g_0\cdot |\{1\leq i\leq n\mathrel{|} \forall \sigma\in\mathfrak S_0(w_\sigma(i)=i \wedge \lambda_\sigma(i)=e)\}|. \end{align*} \end{thm} \begin{proof} Define elements $w'_\sigma\in W$ and $\lambda_\sigma'\in X$ as in Lemma \ref{Lemxprime}. Let $t=(t_{(i,j)})_{(i,j)}\in M^{\mathrm{loc}}(\mathbb{F})$ and let $(t_{i,\pm})_i$ be the two associated tuples of Remark \ref{RemarkLocModDecompUni3}. Let $(i,j)\in\ZZ\times\ZZ$. We have the following equivalences. \begin{align*} \Lambda_{(i,j),\mathbb{F}}=\im \rho_{(i-1,j),+,\mathbb{F}}+t_{(i,j)}&\Leftrightarrow \Lambda_{i,+,\mathbb{F}}=\im \rho_{i-1,+,\mathbb{F}}+t_{i,+},\\ \Lambda_{(i,j),\mathbb{F}}=\im \rho_{(i,j-1),-,\mathbb{F}}+t_{(i,j)}&\Leftrightarrow \Lambda_{j,-,\mathbb{F}}=\im \rho_{j-1,-,\mathbb{F}}+t_{j,-}. \end{align*} Assume now that $t$ lies in the $\Aut(\mathcal L^{\mathrm{inert}})(\mathbb{F})$-orbit corresponding to $x$ under the identifications of Corollary \ref{CorIndexSetUni2}. Consider the chain of neighbors \begin{equation*} \Lambda_{(0,0)}\subset \Lambda_{(1,0)}\subset \dots\subset \Lambda_{(n,0)}\subset \Lambda_{(n,1)}\subset \dots\subset \Lambda_{(n,n)}=\pi_0^{-1}\Lambda_{(0,0)}, \end{equation*} and let $1\leq i\leq n$. By Propositions \ref{ProppRankFirstVersionGen} and \ref{PropEquivCondGen}, the claim of the theorem follows once we can show the following equivalences. \begin{align*} \Lambda_{i,+,\mathbb{F}}=\im \rho_{i-1,+,\mathbb{F}}+t_{i,+}&\Leftrightarrow \forall\sigma\in\mathfrak S_0(w_\sigma(i)=i\wedge \lambda_\sigma(i)=0),\\ \Lambda_{i,-,\mathbb{F}}=\im \rho_{i-1,-,\mathbb{F}}+t_{i,-}&\Leftrightarrow \forall\sigma\in\mathfrak S_0(w'_\sigma(i)=i\wedge \lambda'_\sigma(i)=0). \end{align*} These equivalences follow from Proposition \ref{PropDifferentEmbeddingsIntoFlag2} and Lemma \ref{Lemxprime} by the arguments of the proof of Theorem \ref{ThmPrank}. \end{proof} \subsection{An application to the dimension of the \texorpdfstring{$p$}{p}-rank 0 locus} Assume from now on that $F_0=\QQ$, so that $F/\QQ$ is an imaginary quadratic extension in which $p$ splits. We write $r=r_{\id_\QQ}$ and $s=s_{\id_\QQ}$, so that $n=r+s$. Also write $I_n=\{1,\dots,n\}$. Note that the moduli problem $\mathcal A$ is a special case of the ``fake unitary case'' considered in \cite{haines}. Concretely, the moduli problem defined in \cite[\textsection 5.2]{haines} specializes to $\mathcal A$ for $D=F$. Denote by $\ell:\widetilde{W}\to \NN$ the length function defined in \cite[\textsection 2.1]{gy1}. \begin{lem} Let $x\in\mathrm{Perm}_r$. The smooth $\mathbb{F}$-variety $\mathcal A_{x}$ is equidimensional of dimension $\ell(x)$. \end{lem} \begin{proof} We know from \cite[Lemma 13.1]{haines} that $\mathcal A_{x}$ is non-empty. The rest of the proof is identical to the one of Lemma \ref{LemPropsOfKR}. \end{proof} Let us state Theorem \ref{ThmPrankUni3} in this special case. \begin{thm}\label{ThmPrankSpecialCase} Let $x\in\mathrm{Perm}_r$. Write $x=wu^\lambda$ with $w\in W,\lambda\in X$. Then the $p$-rank on $\mathcal A_x$ is constant with value $|\mathrm{Fix}(w)|$, where $\mathrm{Fix}(w)=\{i\in I_n\mathrel{|} w(i)=i\}$. \end{thm} We want to use this result to compute the dimension of the $p$-rank 0 locus in $\mathcal A_{\mathbb{F}}$. We do this by copying the approach of \cite[\textsection 8]{gy2}. Denote by $\mathrm{Perm}_r^{(0)}$ the subset of those $x\in \mathrm{Perm}_r$ such that the $p$-rank on $\mathcal A_{x}$ is equal to 0. Denote by $W_{n,r}$ the subset of those $w \in W$ satisfying $\mathrm{Fix}(w)=\emptyset$ and \begin{equation}\label{EqDefnWnr} |\{i\in I_n\mathrel{|} w(i)<i\}|=r. \end{equation} \begin{lem}[{Cf.\ \cite[Lemma 8.1]{gy2}}]\label{LemPerm0} The canonical projection $\widetilde{W}\to W$ induces a bijection $\mathrm{Perm}_r^{(0)}\to W_{n,r}$. Its inverse is given by $w\mapsto u^{\lambda(w)}w$ with \begin{equation*} \lambda(w)(i) = \begin{cases} 0, & \text{if } w^{-1}(i) > i \\ 1, & \text{if } w^{-1}(i) < i \end{cases}, \quad i\in I_n. \end{equation*} \end{lem} \begin{proof} This is an easy combinatorial consequence of Theorem \ref{ThmPrankSpecialCase} and the interpretation of $\mathrm{Perm}_r$ in terms of extended alcoves, see Section \ref{SecExtAffWeylUni2}. \end{proof} Define for $\sigma\in S_n$ the following sets and natural numbers. \begin{align*} A_\sigma&=\{(i,j)\in (I_n)^2\mathrel{|} i<j<\sigma(j)<\sigma(i)\},&a_\sigma&=|A_\sigma|,\\ \tilde{A}_\sigma&=\{(i,j)\in (I_n)^2\mathrel{|} \sigma(j)<\sigma(i)<i<j\},&\tilde{a}_\sigma&=|\tilde{A}_\sigma|,\\ B_\sigma&=\{(i,j)\in (I_n)^2\mathrel{|} \sigma(i)<i<j<\sigma(j)\},&b_\sigma&=|B_\sigma|,\\ \tilde{B}_\sigma&=\{(i,j)\in (I_n)^2\mathrel{|} i<\sigma(i)<\sigma(j)<j\},&\tilde{b}_\sigma&=|\tilde{B}_\sigma|,\\ N_\sigma&=a_\sigma+\tilde{a}_\sigma+b_\sigma+\tilde{b}_\sigma. \end{align*} Note that $N_\sigma=N_{\sigma^{-1}}$ in view of the obvious identities $a_\sigma=\tilde{a}_{\sigma^{-1}}$ and $b_\sigma=\tilde{b}_{\sigma^{-1}}$. \begin{prop} Let $u^{\lambda}w\in \mathrm{Perm}_r^{(0)},\ w\in W,\lambda\in X$. Then $\ell(u^{\lambda}w)=N_w$. \end{prop} \begin{proof} Denote by $e_i$ the $i$-th standard basis vector of $\ZZ^n$. The positive roots $\beta>0$ of $\mathrm{GL}_{n}$ are given by $\beta_{ij}=e_i-e_j,\ 1\leq i<j\leq n$. Denote by $\left<\cdot,\cdot\right>$ the standard symmetric pairing on $\ZZ^n$, determined by $\pair{e_i}{e_j}=\delta_{ij}$. By \cite[(8.1)]{gy2} the following Iwahori-Matsumoto formula holds. \begin{equation}\label{EqIwahoriMatsumoto} \ell(u^\lambda w)=\sum_{\substack{\beta>0\\w^{-1}\beta>0}}|\pair{\beta}{\lambda}|+ \sum_{\substack{\beta>0\\w^{-1}\beta<0}}|\pair{\beta}{\lambda}+1|. \end{equation} Using Lemma \ref{LemPerm0}, the equality $\ell(u^{\lambda}w)=N_{w^{-1}}$ readily follows. \end{proof} Define \begin{equation*} N_{n,r}:=\min\bigl((r-1)(n-r),r(n-r-1)\bigr)= \begin{cases} (r-1)(n-r),&\text{if } r\leq n/2,\\ r(n-r-1),&\text{if } r\geq n/2. \end{cases} \end{equation*} \begin{prop} Let $\sigma\in W_{n,r}$. Then $N_\sigma\leq N_{n,r}$. \end{prop} \begin{proof} Consider the set $M=\{(n,r,i_0)\in\NN^3\mathrel{|} 1\leq r\leq n-1,\ 2\leq i_0\leq n\}$ and equip it with the lexicographical ordering $<$, which is a well-ordering on $M$. For $(n,r,i_0)\in M$ we define $W_{n,r,i_0}=\{\sigma\in W_{n,r}\mathrel{|} \min\{2\leq i\leq n\mathrel{|} \sigma(i)<i\}=i_0\}$. Denote by $\mathcal P(n,r,i_0)$ the following statement. \begin{equation*} \forall \sigma\in W_{n,r,i_0}:\ N_\sigma\leq N_{n,r}. \end{equation*} We will prove it by induction on $(n,r,i_0)$. Let $(n,r,i_0)\in M,\ \sigma\in W_{(n,r,i_0)}$ and assume that $\mathcal P(n',r',i'_0)$ is true for all $(n',r',i'_0)\in M$ with $(n',r',i'_0)<(n,r,i_0)$. Set $\sigma'=\sigma\circ (i_0-1,i_0)$. We distinguish the following four cases. \textbf{Case 1:} $\sigma(i_0)<i_0-1$ and $\sigma(i_0-1)>i_0$. \textbf{Case 2:} $\sigma(i_0)=i_0-1$ and $\sigma(i_0-1)>i_0$. \textbf{Case 3:} $\sigma(i_0)<i_0-1$ and $\sigma(i_0-1)=i_0$. \textbf{Case 4:} $\sigma(i_0)=i_0-1$ and $\sigma(i_0-1)=i_0$. We use the example of \textbf{Case 2} to illustrate how to proceed. So assume that $\sigma(i_0)=i_0-1$ and $\sigma(i_0-1)>i_0$. We read off the following identities. \begin{align*} a_\sigma&=a_{\sigma'},\\ \tilde{a}_\sigma&=\tilde{a}_{\sigma'}+|\{j\in I_n\mathrel{|} \sigma(j)<i_0-1<i_0<j\}|,\\ b_\sigma&=b_{\sigma'}+|\{j\in I_n\mathrel{|} i_0-1<i_0<j<\sigma(j)\}|,\\ \tilde{b}_\sigma&=\tilde{b}_{\sigma'}+|\{i\in I_n\mathrel{|} i<\sigma(i)<i_0-1<i_0\}|. \end{align*} Identifying $\{1,\ldots,\widehat{i_0-1},\ldots,n\}$ with $\{1,\ldots,n-1\}$, we consider the restriction $\sigma'\big|_{\{1,\ldots,\widehat{i_0-1},\ldots,n\}}$ as an element of $W_{n-1,r-1,j_0}$ for some $j_0$. By induction hypothesis we know that $N_{\sigma'}=N_{\sigma'|_{\{1,\ldots,\widehat{i_0-1},\ldots,n\}}}\leq N_{n-1,r-1}$. In view of $N_{n,r}-N_{n-1,r-1} \geq n-r-1$ it therefore suffices to show the following two inequalities. \begin{align*} &i_0-2\geq |\{j\in I_n\mathrel{|} \sigma(j)<i_0-1<i_0<j\}|+|\{i\in I_n\mathrel{|} i<\sigma(i)<i_0-1\}|,\\ &n-r-(i_0-1)\geq |\{j\in I_n\mathrel{|} i_0<j<\sigma(j)\}|. \end{align*} For the first inequality, it suffices to note that $\sigma$ maps both sets in question into $I_{i_0-2}$. On the other hand, by the definition of $i_0$ we have $I_{i_0-1}\subset \{i\in I_n\mathrel{|} i<\sigma(i)\}$, so that \eqref{EqDefnWnr} implies the second inequality. \end{proof} \begin{prop} We have \begin{equation*} \max_{\sigma\in W_{n,r}}N_\sigma=N_{n,r}. \end{equation*} \end{prop} \begin{proof} It suffices to show that there is a $\sigma\in W_{n,r}$ satisfying $N_\sigma=N_{n,r}$. As $W_{n,r}\to W_{n,n-r},\ \sigma\mapsto \sigma^{-1}$ is a bijection and as $N_\sigma=N_{\sigma^{-1}}$, we may assume that $r\leq n/2$. One easily checks that \begin{equation*} \sigma=(1,2)(3,4)\cdots(2(r-1)-1,2(r-1))(2r-1,2r,2r+1,\ldots,n)\in W_{n,r} \end{equation*} satisfies $N_\sigma=(r-1)(n-r)=N_{n,r}$. \end{proof} Denote by $\mathcal A^{(0)} \subset \mathcal A(\mathbb{F})$ the subset where the $p$-rank of the underlying abelian variety is equal to $0$. It is a closed subset and we equip it with the reduced scheme structure. From the discussion above we obtain the following result. \begin{thm}\label{ThmPRank0Locus} $\dim \mathcal A^{(0)}=\min\bigl((r-1)(n-r),r(n-r-1)\bigr)$. \end{thm} \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} In recent years quickly growing interest in pricing of credit-risky securities (e.g., defaultable bonds) has been seen in the mathematical finance literature. One of the basic models (for applications see for instance Chen and Joslin \cite{CheJos}) is the following two-dimensional affine diffusion process: \begin{align}\label{DD} \begin{cases} \dd Y_t=(a-bY_t)\,\dd t+\sqrt{Y_t}\,\dd W_t,\\ \dd X_t= (m-\theta X_t)\,\dd t+\sqrt{Y_t}\,\dd B_t, \end{cases} \quad t\geq 0, \end{align} where \ $a$, $b$, $\theta$ \ and \ $m$ \ are real parameters such that \ $a>0$ \ and \ $B$ \ and \ $W$ \ are independent standard Wiener processes. Note that \ $Y$ \ is a Cox-Ingersol-Ross (CIR) process. For practical use, it is important to estimate the appearing parameters from some discretely observed real data set. In the case of the one-dimensional CIR process, the parameter estimation of \ $a$ \ and \ $b$ \ goes back to Overbeck and Ryd\'en \cite{OveRyd}, Overbeck \cite{Ove}, and see also the very recent papers of Ben Alaya and Kebaier \cite{BenKeb1,BenKeb2}. For asymptotic results on discrete time critical branching processes with immigration, one may refer to Wei and Winnicki \cite{WeiWin1} and \cite{WeiWin2}. The process \ $(Y,X)$ \ given by \eqref{DD} is a very special affine process. The set of affine processes contains a large class of important Markov processes such as continuous state branching processes and Orstein-Uhlenbeck processes. Further, a lot of models in financial mathematics are also special affine processes such as the Heston model \cite{Hes}, the model due to Barndorff-Nielsen and Shephard \cite{BarShe} or the model due to Carr and Wu \cite{CarWu}. A precise mathematical formulation and complete characterization of regular affine processes are due to Duffie et al. \cite{DufFilSch}. Later several authors have contributed to the study of properties of general affine processes: to name a few, Andersen and Piterbarg \cite{AndPit} (moment explosions in stochastic volatility models), Dawson and Li \cite{DawLi} (jump-type SDE representation for two-dimensional affine processes), Filipovi\'{c} and Mayerhofer \cite{FilMay} (applications to the pricing of bond and stock options), Glasserman and Kim \cite{GlaKim} (the range of finite exponential moments and the convergence to stationarity in affine diffusion models), Jena et al. \cite{JenKimXin} (long-term and blow-up behaviors of exponential moments in multi-dimensional affine diffusions), Keller-Ressel et al. \cite{KelSchTei1,KelSchTei2} (stochastically continuous, time-homogeneous affine processes with state space \ $\RR_+^n\times\RR^d$ \ or more general ones are regular). We also refer to the overview articles Cuchiero et al. \cite{CucFilTei} and Friz and Keller-Ressel \cite{FriKel}. To the best knowledge of the authors the parameter estimation problem for multi-dimensional affine processes has not been tackled so far. Since affine processes are being used in financial mathematics very frequently, the question of parameter estimation for them is of high importance. Our aim is to start the discussion with a simple non-trivial example: the two-dimensional affine diffusion process given by \eqref{DD}. The article is divided into two parts and there are two appendices. In Section~\ref{Section_scaling} we recall some notations, the definition of affine processes and some of their basic properties, and then a simple set of sufficient conditions for the weak convergence of scaled affine processes is presented. Roughly speaking, given a family of affine processes \ $(Y^{(\theta)}(t), X^{(\theta)}(t))_{t\geq 0}$, \ $\theta>0$, \ such that the corresponding admissible parameters converge in an appropriate way (see Theorem \ref{Thm1}), the scaled process \ $\left( \theta^{-1} Y^{(\theta)}(\theta t), \theta^{-1} X^{(\theta)}(\theta t) \right)_{t\geq 0}$ \ converge weakly towards an affine diffusion process as \ $\theta \to \infty$. \ We specialize our result for one-dimensional continuous state branching processes with immigration which generalizes Theorem 2.3 in Huang et al.\ \cite{HuaMaZhu}. The scaling Theorem \ref{Thm1} is proved for quite general affine processes since it might have applications elsewhere later on. In Section~\ref{Section_statistics} the scaling Theorem \ref{Thm1} is applied to study the asymptotic behavior of least squares and conditional least squares estimators of some parameters of a critical two-dimensional affine diffusion process given by \eqref{DD}, see Theorems \ref{Thm2}, \ref{Thm3} and \ref{Thm4}. In Appendix \ref{AppendixA} we check that some integrals in the form of the infinitesimal generator of an affine process that we use are well-defined. Appendix \ref{AppendixB} is devoted to show that the least squares estimator of \ $m$ \ cannot be asymptotically weakly consistent. \section{A scaling theorem for affine processes}\label{Section_scaling} Let \ $\NN$, \ $\ZZ_+$, \ $\RR$, \ $\RR_+$, \ $\RR_-$, \ $\RR_{++}$, \ and \ $\CC$ \ denote the sets of positive integers, non-negative integers, real numbers, non-negative real numbers, non-positive real numbers, positive real numbers and complex numbers, respectively. For \ $x , y \in \RR$, \ we will use the notations \ $x \land y := \min(x, y)$ \ and \ $x \vee y := \max(x, y)$. \ For \ $x, y\in \CC^k$, $k\in\NN$, \ we write \ $\langle x,y\rangle := \sum_{i=1}^kx_iy_i$ \ (notice that this is not the scalar product on \ $\CC^k$, \ however for \ $x\in\CC^k$ \ and \ $y\in\RR^k$, \ $\langle x,y\rangle$ \ coincides with the usual scalar product of \ $x$ \ and \ $y$). \ By \ $\|x\|$ \ and \ $\|A\|$ \ we denote the Euclidean norm of a vector \ $x\in\RR^p$ \ and the induced matrix norm of a matrix \ $A\in\RR^{p\times p}$, \ respectively. Further, let \ $U:=\{z_1+iz_2 : z_1\in\RR_-, \, z_2\in\RR\}\times(i\RR^d)$. \ By \ $C^2_c(\RR_+\times\RR^d)$ \ ($C^\infty_c(\RR_+\times\RR^d)$) \ we denote the set of twice (infinitely) continuously differentiable complex-valued functions on \ $\RR_+\times\RR^d$ \ with compact support, where \ $d\in\NN$. \ The set of c\`{a}dl\`{a}g functions from \ $\RR_+$ \ to \ $\RR_+\times\RR^d$ \ will be denoted by \ $\DD(\RR_+,\RR_+\times\RR^d)$. \ For a bounded function \ $g:\RR_+\times\RR^d\to\RR^p$, \ let \ $\|g\|_\infty := \sup_{x\in\RR_+\times\RR^d} \|g(x)\|$. \ Convergence in distribution, in probability and almost sure convergence will be denoted by \ $\distr$, \ $\stoch$ \ and \ $\as$, \ respectively. Next we briefly recall the definition of affine processes with state space \ $\RR_+\times\RR^d$ \ based on Duffie et al.\ \cite{DufFilSch}. \begin{Def} A transition semigroup \ $(P_t)_{t\in\RR_+}$ \ with state space \ $\RR_+\times\RR^d$ \ is called a (general) affine semigroup if its characteristic function has the representation \begin{align}\label{affine_SG} \int_{\RR_+\times\RR^d} \ee^{\langle u,\xi\rangle} P_t(x,\dd\xi) = \ee^{\langle x,\psi(t,u)\rangle + \phi(t,u)} \end{align} for \ $x\in \RR_+\times\RR^d$, \ $u\in U$ \ and \ $t\in\RR_+$, \ where \ $\psi(t,\cdot)=(\psi_1(t,\cdot),\psi_2(t,\cdot))\in\CC\times\CC^d$ \ is a continuous \ $\CC^{1+d}$-valued function on \ $U$ \ and \ $\phi(t,\cdot)$ \ is a continuous \ $\CC$-valued function on \ $U$ \ satisfying \ $\phi(t,0)=0$. \ The affine semigroup \ $(P_t)_{t\in\RR_+}$ \ defined by \eqref{affine_SG} is called regular if it is stochastically continuous (equivalently, for all \ $u\in U$, \ the functions \ $\RR_+\ni t\mapsto \Psi(t,u)$ \ and \ $\RR_+\ni t\mapsto \phi(t,u)$ \ are continuous) and \ $\partial_1\psi(0,u)$ \ and \ $\partial_1\phi(0,u)$ \ exist for all \ $u\in U$ \ and are continuous at \ $u=0$ \ (where \ $\partial_1\psi$ \ and \ $\partial_1\phi$ \ denote the partial derivatives of \ $\psi$ \ and \ $\phi$, \ respectively, with respect to the first variable). \end{Def} \begin{Rem} We call the attention that Duffie et al. \cite{DufFilSch} in their Definition 2.1 assume only that Equation \eqref{affine_SG} hold for \ $x\in \RR_+\times\RR^d$, $u\in \partial U=i\RR^{1+d}$, $t\in\RR_+$, \ i.e., instead of \ $u\in U$ \ they only require that \ $u$ \ should be an element of the boundary \ $\partial U$ \ of \ $U$. \ However, by Proposition 6.4 in Duffie et al. \cite{DufFilSch}, one can formulate the definition of a regular affine process as we did. Note also that this kind of definition was already given by Dawson and Li \cite[Definitions 2.1 and 3.3]{DawLi}. Finally, we remark that every stochastically continuous affine semigroup is regular due to Keller-Ressel et al. \cite[Theorem 5.1]{KelSchTei1}.\proofend \end{Rem} \begin{Def}\label{Def_admissible} A set of parameters \ $( a, \alpha, b, \beta, m, \mu)$ \ is called admissible if \renewcommand{\labelenumi}{{\rm(\roman{enumi})}} \begin{enumerate} \item $a = (a_{i,j})_{i,j=1}^{1+d} \in \RR^{(1+d)\times(1+d)}$ \ is a symmetric positive semidefinite matrix with \ $a_{1,1}=0$ \ (hence \ $a_{1,k} = a_{k,1} = 0$ \ for all \ $k \in \{2,\ldots,1+d\}$), \item $\alpha = (\alpha_{i,j})_{i,j=1}^{1+d} \in \RR^{(1+d)\times(1+d)}$ \ is a symmetric positive semidefinite matrix, \item $b = (b_i)_{i=1}^{1+d} \in \RR_+\times\RR^d$, \item $\beta = (\beta_{i,j})_{i,j=1}^{1+d} \in \RR^{(1+d)\times(1+d)}$ \ with \ $\beta_{1,j}=0$ \ for all \ $j \in \{2,\ldots,1+d\}$, \item $m(\dd\xi) = m(\dd\xi_1, \dd\xi_2)$ \ is a \ $\sigma$-finite measure on \ $\RR_+\times\RR^d$ \ supported by \ $(\RR_+\times\RR^d)\setminus\{(0,0)\}$ \ such that \[ \int_{\RR_+\times\RR^d} \left[ \xi_1 + (\|\xi_2\| \land \|\xi_2\|^2) \right] m(\dd\xi) <\infty, \] \item $\mu(\dd\xi) = \mu(\dd\xi_1, \dd\xi_2)$ \ is a $\sigma$-finite measure on \ $\RR_+\times\RR^d$ \ supported by \ $(\RR_+\times\RR^d)\setminus\{(0,0)\}$ \ such that \[ \int_{\RR_+\times\RR^d} \|\xi\| \land \|\xi\|^2 \mu(\dd\xi) <\infty. \] \end{enumerate} \end{Def} \begin{Rem} Note that our Definition \ref{Def_admissible} of the set of admissible parameters is not so general as Definition 2.6 in Duffie et al. \cite{DufFilSch}. Firstly, the set of admissible parameters is defined only for affine process with state space \ $\RR_+\times\RR^d$, \ while Duffie et al. \cite{DufFilSch} consider affine processes with state space \ $\RR_+^n\times\RR^d$. We restrict ourselves to this special case, since our scaling Theorem \ref{Thm1} is valid only in this case. Secondly, our conditions (v) and (vi) of Definition \ref{Def_admissible} are stronger than that of (2.10) and (2.11) of Definition 2.6 in Duffie et al. \cite{DufFilSch}. Thirdly, according to our definition, a set of admissible parameters does not contain parameters corresponding to killing, while in Definition 2.6 in Duffie et al. \cite{DufFilSch} such parameters are included. Our definition of admissible parameters can be considered as a \ $(1+d)$-dimensional version of Definition 6.1 in Dawson and Li \cite{DawLi}. The reason for this definition is to have a more pleasant form of the infinitesimal generator of an affine process compared to that of Duffie et al. \cite[formula (2.12)]{DufFilSch}. For more details, see Remark \ref{Rem4}.\proofend \end{Rem} \begin{Thm}{(Duffie et al. \cite[Theorem 2.7]{DufFilSch})}\label{Thm_Duffie_et_al} Let \ $(a, \alpha, b, \beta, m, \mu)$ \ be a set of admissible parameters. Then there exists a unique regular affine semigroup \ $(P_t)_{t\in\RR_+}$ \ with infinitesimal generator \begin{align}\label{affine_inf_gen} \begin{split} (\cA f)(x) &= \sum_{i,j=1}^{1+d} (a_{i,j} + \alpha_{i,j} x_1) f_{i,j}''(x) + \langle f'(x), b + \beta x \rangle \\ &\phantom{\quad} + \int_{\RR_+\times \RR^d} (f(x+\xi) - f(x) - \langle f'_{(2)}(x), \xi_2 \rangle) \, m(\dd\xi)\\ &\phantom{\quad} + \int_{\RR_+\times \RR^d} (f(x+\xi) - f(x) - \langle f'(x), \xi \rangle) x_1 \, \mu(\dd\xi) \end{split} \end{align} for \ $x=(x_1,x_2)\in\RR_+\times\RR^d$ \ and \ $f\in C^2_c(\RR_+\times\RR^d)$, \ where \ $f_i'$, \ $i\in\{1,\ldots,1+d\}$, \ and \ $f_{i,j}''$, \ $i,j\in\{1,\ldots,1+d\}$, \ denote the first and second order partial derivatives of $f$ with respect to its \ $i$-th and \ $i$-th and \ $j$-th variables, and \ $f'(x) := (f_1'(x),\ldots,f_{1+d}'(x))^\top$, \ $f'_{(2)}(x) := (f_2'(x),\ldots,f_{1+d}'(x))^\top$. Further, \ $\cC_c^\infty(\RR_+\times\RR^d)$ \ is a core of \ $\cA$. \end{Thm} \begin{Rem}\label{Rem4} Note that the form of the infinitesimal generator \ $\cA$ \ in Theorem \ref{Thm_Duffie_et_al} is slightly different from the one given in (2.12) in Duffie et al. \cite{DufFilSch}. Our formula \eqref{affine_inf_gen} is in the spirit of Dawson and Li \cite[formula (6.5)]{DawLi}. On the one hand, the point is that under the conditions (v) and (vi) of Definition \ref{Def_admissible}, one can rewrite (2.12) in Duffie et al. \cite{DufFilSch} into the form \eqref{affine_inf_gen}, by changing the \ $2$-nd, $\ldots$, $(1+d)$-th coordinates of \ $b\in\RR_+\times\RR^d$ \ and the first column of \ $\beta\in\RR^{(1+d)\times(1+d)}$, \ respectively, in appropriate ways (see Appendix \ref{AppendixA}). To see this, it is enough to check that the integrals in \eqref{affine_inf_gen} are well-defined (i.e., elements of \ $\CC$) \ under the conditions (v) and (vi) of Definition \ref{Def_admissible}. For further details, see also Appendix~\ref{AppendixA}. On the other hand, the killing rate (see page 995 in Duffie et al. \cite{DufFilSch}) of the affine semigroup \ $(P_t)_{t\in\RR_+}$ \ in Theorem \ref{Thm_Duffie_et_al} is identically zero. This also implies that the affine processes that we will consider later on will have lifetime infinity.\proofend\looseness=-1 \end{Rem} \begin{Rem} In dimension 2 (i.e., if \ $d=1$), by Theorem 6.2 in Dawson and Li \cite{DawLi} and Theorem 2.7 in Duffie et al. \cite{DufFilSch} (see also Theorem \ref{Thm_Duffie_et_al}), for an infinitesimal generator \ $\cA$ \ given by \eqref{affine_inf_gen} with \ $d=1$ \ one can construct a two-dimensional system of jump type SDEs of which there exists a pathwise unique strong solution \ $(Y(t),X(t))_{t\in\RR_+}$ \ which is a regular affine Markov process with the given infinitesimal generator \ $\cA$.\proofend \end{Rem} The next lemma is simple but very useful. \begin{Lem}\label{Lem_inf_gen} Let \ $(Z(t))_{t\in\RR_+}$ \ be a time-homogeneous Markov process with state space \ $\RR_+\times\RR^d$ \ and let us denote its infinitesimal generator by \ $\cA_Z$. \ Suppose that \ $\cC_c^2(\RR_+\times\RR^d)$ \ is a subset of the domain of \ $\cA_Z$. \ Then for all \ $\theta\in\RR_{++}$, \ the time-homogeneous Markov process \ $(Z_\theta(t))_{t\in\RR_+} := (\theta^{-1}Z(\theta t))_{t\in\RR_+}$ \ has infinitesimal generator \begin{align*} (\cA_{Z_\theta}f)(x) = \theta (\cA_Z f_\theta)(\theta x), \qquad x\in \RR_+\times\RR^d, \quad f\in C_c^2(\RR_+\times\RR^d), \end{align*} where \ $f_\theta(x) := f(\theta^{-1}x)$, \ $x\in \RR_+\times\RR^d$. \end{Lem} \noindent{\bf Proof.} By definition, the infinitesimal generator of \ $(Z_\theta(t))_{t\in\RR_+}$ \ takes the form \begin{align*} (\cA_{Z_\theta}f)(x) &= \lim_{t\downarrow 0} \frac{\EE( f(\theta^{-1} Z(\theta t)) \mid \theta^{-1}Z(0)=x) - f(x)}{t}\\ &= \lim_{t\downarrow 0} \frac{\theta \EE( f_\theta(Z(\theta t))\mid Z(0)=\theta x) - \theta f_\theta(\theta x)} {\theta t}\\ &= \theta \lim_{t'\downarrow 0} \frac{\EE( f_\theta(Z(t')) \mid Z(0)=\theta x) - f_\theta(\theta x)}{t'} = \theta (\cA_Z f_\theta)(\theta x) \end{align*} for all \ $x\in \RR_+\times\RR^d$ \ and \ $f\in C_c^2(\RR_+\times\RR^d)$. \proofend \begin{Thm}\label{Thm1} For all \ $\theta\in\RR_{++}$, \ let \ $(Y^{(\theta)}(t), X^{(\theta)}(t))_{t\in\RR_+}$ \ be a \ $(1+d)$-dimensional affine process with state space \ $\RR_+\times\RR^d$ \ and with admissible parameters \ $(a^{(\theta)}, \alpha^{(\theta)}, b^{(\theta)}, \beta^{(\theta)}, m, \mu)$ \ such that additionally \begin{align}\label{cond1} \int_{\RR_+\times\RR^d} \|\xi\| \, m(\dd\xi)<\infty \qquad\text{and}\qquad \int_{\RR_+\times\RR^d} \|\xi\|^2 \, \mu(\dd\xi)<\infty. \end{align} Let \ $a, \alpha, \beta \in \RR^{(1+d)\times(1+d)}$, \ $b\in\RR_+\times\RR^d$, \ and let \ $(Y(t), X(t))_{t\in\RR_+}$ \ be a \ $(1+d)$-dimensional affine process with state space \ $\RR_+\times\RR^d$ \ and with the set of admissible parameters \ $(a, \widetilde{\alpha}, \widetilde{b}, \beta, 0, 0)$, \ where \[ \widetilde{\alpha} := \alpha + \frac{1}{2} \int_{\RR_+\times\RR^d} \xi \xi^\top \, \mu(\dd\xi) , \] and \ $\widetilde{b} = (\widetilde{b}_i)_{i=1}^{1+d}$ \ with \ $\widetilde{b}_i := b_i$ \ for \ $i \in \{2,\ldots,1+d\}$ \ and \[ \widetilde{b}_1 := b_1 + \int_{\RR_+\times\RR^d} \xi_1 \, m(\dd\xi) . \] If \begin{gather*} \theta^{-1} a^{(\theta)} \to a, \qquad \alpha^{(\theta)} \to \alpha, \qquad b^{(\theta)} \to b, \qquad \theta \beta^{(\theta)} \to \beta, \\ \theta^{-1}(Y^{(\theta)}(0), X^{(\theta)}(0)) \distr (Y(0), X(0)) \end{gather*} as \ $\theta\to\infty$, \ then \begin{align*} \left( Y^{(\theta)}_\theta(t), X^{(\theta)}_\theta(t)\right)_{t\in\RR_+} = \left(\theta^{-1}Y^{(\theta)}(\theta t), \theta^{-1}X^{(\theta)}(\theta t)\right)_{t\in\RR_+} \distr (Y(t),X(t))_{t\in\RR_+} \end{align*} in \ $\DD(\RR_+,\RR_+\times\RR^d)$ \ as \ $\theta\to\infty$. \end{Thm} \begin{Rem} (i) Note that the limit process \ $(Y(t), X(t))_{t\in\RR_+}$ \ in Theorem \ref{Thm1} has continuous sample paths almost surely. However, this is not a big surprise, since in condition \eqref{cond1} of Theorem \ref{Thm1} we require finite second moment for the measure \ $\mu$. \noindent (ii) Note also that the matrix \ $\widetilde{\alpha}\in\RR^{(1+d)\times(1+d)}$ \ given in Theorem \ref{Thm1} is symmetric and positive semidefinite, since \ $\alpha$ \ is symmetric and positive semidefinite, and for all \ $z\in\RR^{1+d}$, \begin{align*} \left\langle \int_{\RR_+\times\RR^d} \xi \xi^\top \, \mu(\dd\xi) z,z \right\rangle = \int_{\RR_+\times\RR^d} (z^\top\xi)^2 \, \mu(\dd\xi) \geq 0.\\[-26pt] \end{align*} \proofend \end{Rem} \noindent{\bf Proof of Theorem \ref{Thm1}.} By Duffie et al. \cite[Theorem 2.7]{DufFilSch}, \ $\cC_c^\infty(\RR_+\times\RR^d)$ \ is a core of the infinitesimal generator \ $\cA_{(Y,X)}$ \ of the process \ $(Y(t), X(t))_{t\in\RR_+}$, \ and hence \ $\{(f,\cA_{(Y,X)}f) : f\in\cD(\cA_{(Y,X)})\}$ \ coincides with the closure of \ $\{(f,\cA_{(Y,X)}f) : f\in\cC_c^\infty(\RR_+\times\RR^d)\}$, \ where \ $\cD(\cA_{(Y,X)})$ \ denotes the domain of \ $\cA_{(Y,X)}$, \ see, e.g., Ethier and Kurtz \cite[page 17]{EthKur}. In other words, the closure of \ $\{(f,\cA_{(Y,X)}f) : f\in\cC_c^\infty(\RR_+\times\RR^d)\}$ \ generates the affine semigroup corresponding to \ $\cA_{(Y,X)}$. Next we show that for all \ $f\in \cC_c^\infty(\RR_+\times\RR^d)$, \ we have \begin{align}\label{help2} \lim_{\theta\to\infty} \sup_{x\in\RR_+\times\RR^d} \Big| (\cA_{(Y_\theta^{(\theta)},X_\theta^{(\theta)})}f)(x) - (\cA_{(Y,X)} f)(x) \Big| =0. \end{align} First note that it is enough to prove \eqref{help2} for real-valued functions \ $f\in \cC_c^\infty(\RR_+\times\RR^d)$, \ since if \eqref{help2} holds for for real-valued functions \ $f\in \cC_c^\infty(\RR_+\times\RR^d)$, \ then, by decomposing \ $f$ \ into real and imaginary parts, the linearity of the infinitesimal generators in question and triangular inequality yield \eqref{help2} for complex-valued functions \ $f\in \cC_c^\infty(\RR_+\times\RR^d)$. Hence in what follows without loss of generality we can assume that \ $f\in \cC_c^\infty(\RR_+\times\RR^d)$ \ is real-valued. \allowdisplaybreaks For all \ $f\in\cC_c^\infty(\RR_+\times\RR^d)$, \ $\theta\in\RR_{++}$, \ and \ $x\in \RR_+\times\RR^d$, \ we have \begin{align}\label{help1} \begin{split} & f_\theta(x) = f(\theta^{-1}x),\\ & (f_\theta)_i'(x) = \theta^{-1} f_i'(\theta^{-1}x),\qquad i\in\{1,\ldots,1+d\},\\ & (f_\theta)_{i,j}'' (x) = \theta^{-2} f_{i,j}'' (\theta^{-1}x),\qquad i,j\in\{1,\ldots,1+d\}. \end{split} \end{align} Then, by Lemma \ref{Lem_inf_gen}, \eqref{affine_inf_gen} and \eqref{help1}, \begin{align*} (\cA_{(Y_\theta^{(\theta)}, X_\theta^{(\theta)})}f)(x) &= \theta (\cA_{(Y^{(\theta)}, X^{(\theta)})}f_\theta)(\theta x)\\ &=\theta \Biggl[ \sum_{i,j=1}^{1+d} (a_{i,j}^{(\theta)} + \alpha_{i,j}^{(\theta)} \theta x_1) \theta^{-2} f_{i,j}''(\theta^{-1} \theta x) + \langle \theta^{-1} f'(\theta^{-1} \theta x), b^{(\theta)} + \beta^{(\theta)} \theta x \rangle \\ &\quad + \int_{\RR_+\times \RR^d} (f(\theta^{-1}(\theta x+\xi)) - f(\theta^{-1} \theta x) - \langle \theta^{-1} f'_{(2)}(\theta^{-1} \theta x), \xi_2 \rangle) \, m(\dd\xi) \\ &\quad + \int_{\RR_+\times \RR^d} (f(\theta^{-1}(\theta x+\xi)) - f(\theta^{-1} \theta x) - \langle \theta^{-1} f'(\theta^{-1} \theta x), \xi \rangle) \theta x_1 \, \mu(\dd\xi) \Biggr]\\ &=\sum_{i,j=1}^{1+d} (\theta^{-1} a_{i,j}^{(\theta)} + \alpha_{i,j}^{(\theta)} x_1) f_{i,j}''(x) + \langle f'(x), b^{(\theta)} + \theta \beta^{(\theta)} x \rangle \\ &\phantom{\quad} + \int_{\RR_+\times \RR^d} (f(x+\theta^{-1}\xi) - f(x) - \langle f'_{(2)}(x), \theta^{-1} \xi_2 \rangle) \theta \, m(\dd\xi) \\ &\phantom{\quad} + x_1 \int_{\RR_+\times \RR^d} (f(x+\theta^{-1}\xi) - f(x) - \langle f'(x), \theta^{-1} \xi \rangle) \theta^2 \, \mu(\dd\xi) \end{align*} for \ $f\in\cC_c^\infty(\RR_+\times\RR^d)$ \ and \ $x\in\RR_+\times\RR^d$. \ Hence, for all \ $x=(x_1,x_2)\in\RR_+\times\RR^d$, \ using the triangular inequality and that \ $\vert \langle u,v\rangle\vert \leq \Vert u\Vert \Vert v\Vert$, $u,v\in\RR^p$, \ we have \begin{align*} &\Bigl| (\cA_{(Y_\theta^{(\theta)},X_\theta^{(\theta)})}f)(x) - (\cA_{(Y,X)} f)(x) \Bigr| \\ &\leq \sum_{i,j=1}^{1+d} ( |\theta^{-1} a_{i,j}^{(\theta)} - a_{i,j}| + |\alpha_{i,j}^{(\theta)} - \alpha_{i,j}| x_1 ) |f_{i,j}''(x)|\\ &\phantom{\quad} + ( \|b^{(\theta)} - b\| + \|\theta \beta^{(\theta)} - \beta\| \|x\| ) \|f'(x)\| \\ &\phantom{\quad} + \left| \int_{\RR_+\times \RR^d} \Big( f(x+\theta^{-1}\xi) - f(x) - \theta^{-1} \langle f'(x), \xi \rangle \Big) \theta \, m(\dd\xi) \right| \\ &\phantom{\quad} + x_1 \Bigg| \int_{\RR_+\times \RR^d} \Big( f(x+\theta^{-1}\xi) - f(x) - \theta^{-1} \langle f'(x), \xi \rangle - \frac{1}{2} \theta^{-2} \langle f''(x) \xi, \xi \rangle \Big) \theta^2 \, \mu(\dd\xi) \Bigg| , \end{align*} where \[ f''(x) := \begin{bmatrix} f_{1,1}''(x) & \cdots & f_{1,1+d}''(x) \\ \vdots & \ddots & \vdots \\ f_{1+d,1}''(x) & \cdots & f_{1+d,1+d}''(x) \end{bmatrix} . \] Since \ $f\in\cC_c^\infty(\RR_+\times\RR^d)$, \ we have \begin{gather*} \sup_{x\in\RR_+\times\RR^d} x_1 |f_{i,j}''(x)| < \infty, \qquad \sup_{x\in\RR_+\times\RR^d} |f_{i,j}''(x)| < \infty, \qquad \forall\; i,j\in\{1,\ldots,1+d\}, \\ \sup_{x\in\RR_+\times\RR^d} \|x\| \| f'(x) \| < \infty, \qquad \sup_{x\in\RR_+\times\RR^d} \| f'(x) \| < \infty, \end{gather*} and hence, by our assumptions, in order to prove \eqref{help2} it is enough to check that \begin{align}\label{help3} \lim_{\theta\to\infty} \sup_{x\in\RR_+\times\RR^d} \left| \int_{\RR_+\times \RR^d} \Big( f(x+\theta^{-1}\xi) - f(x) - \theta^{-1} \langle f'(x), \xi \rangle \Big) \theta \, m(\dd\xi) \right| =0, \end{align} and \begin{align}\label{help4} \begin{split} \lim_{\theta\to\infty} \sup_{x\in\RR_+\times\RR^d} x_1 & \left| \int_{\RR_+\times \RR^d} \Big( f(x+\theta^{-1}\xi) - f(x) - \theta^{-1} \langle f'(x), \xi \rangle\right.\\ &\;\left. - \frac{1}{2} \theta^{-2} \langle f''(x) \xi, \xi \rangle \Big) \theta^2 \, \mu(\dd\xi) \right| =0. \end{split} \end{align} First we consider \eqref{help3}. Let \ $\varepsilon\in\RR_{++}$ \ be fixed. Let us choose an \ $M\in\RR_{++}$ \ such that \begin{align}\label{help6} 2 \|f'\|_\infty \int_{(\RR_+\times\RR^d)\setminus ([0,M]\times [-M,M]^d)} \|\xi\| \, m(\dd\xi) < \frac{\varepsilon}{2}. \end{align} In what follows, for abbreviation, \ $[0,M] \times [-M,M]^d$ \ will be denoted by \ $D_M$. \ Such an \ $M$ \ can be chosen, since \ $f\in\cC_c^\infty(\RR_+\times\RR^d)$ \ yields \ $\|f'\|_\infty<\infty$ \ and, by assumption \eqref{cond1}, \ $\int_{\RR_+\times\RR^d} \|\xi\| \, m(\dd\xi)<\infty$. \ By Taylor's theorem, for all \ $\theta\in\RR_{++}$, \ $x\in\RR_+\times\RR^d$ \ and \ $\xi\in D_M$ \ there exists some \ $\tau = \tau(\theta, x, \xi) \in[0,1]$ \ such that \begin{align}\label{help15} f(x+\theta^{-1}\xi) - f(x) = \langle f'(x+\theta^{-1}\tau\xi), \theta^{-1}\xi \rangle. \end{align} Then \begin{multline*} \left| \int_{\RR_+\times \RR^d} \Bigl(f(x+\theta^{-1}\xi) - f(x) - \theta^{-1} \langle f'(x), \xi \rangle \Bigr) \theta \, m(\dd\xi) \right| \\ \leq \int_{\RR_+\times \RR^d} \bigl| \langle f'(x+\theta^{-1}\tau\xi), \xi \rangle - \langle f'(x), \xi \rangle \bigr| \, m(\dd\xi) \leq A^{(1)}_{\theta,M}(x) + A^{(2)}_{\theta,M}(x) \end{multline*} for all \ $x\in\RR_+\times\RR^d$, \ where \begin{align*} &A^{(1)}_{\theta,M}(x) := \int_{D_M} | \langle f'(x+\theta^{-1}\tau\xi) - f'(x), \xi \rangle | \, m(\dd\xi) , \\ &A^{(2)}_{\theta,M}(x) := \int_{(\RR_+\times\RR^d)\setminus D_M} \left( |\langle f'(x+\theta^{-1}\tau\xi), \xi \rangle| + |\langle f'(x), \xi \rangle| \right) m(\dd\xi) . \end{align*} Here \begin{align*} A^{(1)}_{\theta,M}(x) &\leq \int_{D_M} \|f'(x+\theta^{-1}\tau\xi) - f'(x)\| \|\xi\| \, m(\dd\xi) \\ &\leq \sup_{\xi\in D_M} \|f'(x+\theta^{-1}\tau\xi) - f'(x)\| \int_{\RR_+\times\RR^d} \|\xi\| \, m(\dd\xi) . \end{align*} The convexity of \ $D_M$ \ implies \ $\tau \xi = \tau(\theta, x, \xi) \xi \in D_M$ \ for all \ $\theta \in \RR_{++}$, \ $x \in \RR_+\times\RR^d$ \ and \ $\xi \in D_M$, \ and, hence, \[ \sup_{\xi\in D_M} \|f'(x+\theta^{-1}\tau\xi) - f'(x)\| \leq \sup_{\widetilde\xi\in D_M} \|f'(x+\theta^{-1}\widetilde\xi) - f'(x)\| . \] Since \ $f'$ \ is uniformly continuous on \ $\RR_+\times\RR^d$ \ (which follows by mean value theorem using also that \ $\Vert f''\Vert_\infty<\infty$), \ there exists a \ $\theta_0\in\RR_{++}$ \ (depending on \ $\varepsilon$ \ and \ $M$) \ such that \[ \sup_{\widetilde\xi\in D_M} \|f'(x+\theta^{-1}\widetilde\xi) - f'(x)\| \int_{\RR_+\times\RR^d} \|\xi\| \, m(\dd\xi) < \frac{\varepsilon}{2} \] for all \ $x\in\RR_+\times\RR^d$ \ and \ $\theta \in [\theta_0,\infty)$. \ Further, by \eqref{help6}, we have \begin{align*} A^{(2)}_{\theta,M}(x) \leq 2 \|f'\|_\infty \int_{(\RR_+\times\RR^d)\setminus D_M} \|\xi\| \, m(\dd\xi) < \frac{\varepsilon}{2} \end{align*} for all \ $x\in\RR_+\times\RR^d$ \ and \ $\theta\in\RR_{++}$. \ Putting the pieces together we have \eqref{help3}. Now we turn to prove \eqref{help4} in a similar way. Let \ $\varepsilon\in\RR_{++}$ \ be fixed again. Let us now choose an \ $M\in\RR_{++}$ \ such that \begin{align}\label{2help6} 2 \sup_{x \in \RR_+\times\RR^d} x_1 \|f''(x)\| \int_{(\RR_+\times\RR^d)\setminus D_M} \|\xi\|^2 \, \mu(\dd\xi) < \frac{\varepsilon}{2}. \end{align} Such an $M$ can be chosen, since $\sup_{x \in \RR_+\times\RR^d} x_1 \|f''(x)\|<\infty$ for all $f\in\cC_c^\infty(\RR_+\times\RR^d)$ and, by assumption \eqref{cond1}, \ $\int_{\RR_+\times\RR^d} \|\xi\|^2 \, \mu(\dd\xi)<\infty$. \ By Taylor's theorem, for all \ $\theta\in\RR_{++}$, \ $x\in\RR_+\times\RR^d$ \ and \ $\xi\in D_M$ \ there exists some \ $\tau = \tau(\theta, x, \xi) \in[0,1]$ \ such that \begin{align}\label{2help15} f(x+\theta^{-1}\xi) - f(x) - \langle f'(x), \theta^{-1}\xi \rangle = \frac{1}{2} \langle f''(x+\theta^{-1}\tau\xi) \theta^{-1}\xi, \theta^{-1}\xi \rangle. \end{align} Then \begin{align*} &x_1 \left| \int_{\RR_+\times \RR^d} \Big(f(x+\theta^{-1}\xi) - f(x) - \theta^{-1} \langle f'(x), \xi \rangle - \frac{1}{2} \theta^{-2} \langle f''(x) \xi , \xi \rangle \Big) \theta^2 \, \mu(\dd\xi) \right| \\ &\quad\leq \frac{1}{2} x_1 \int_{\RR_+\times \RR^d} \bigl| \langle f''(x+\theta^{-1}\tau\xi) \xi, \xi \rangle - \langle f''(x) \xi, \xi \rangle \bigr| \mu(\dd\xi)\\ &\quad\leq \frac{1}{2} ( B^{(1)}_{\theta,M}(x) + B^{(2)}_{\theta,M}(x) ) \end{align*} for all \ $x\in\RR_+\times\RR^d$, \ where \begin{align*} &B^{(1)}_{\theta,M}(x) := x_1 \int_{D_M} | \langle (f''(x+\theta^{-1}\tau\xi) - f''(x)) \xi, \xi \rangle | \, \mu(\dd\xi) , \\ &B^{(2)}_{\theta,M}(x) := x_1 \int_{(\RR_+\times\RR^d)\setminus D_M} \left( |\langle f''(x+\theta^{-1}\tau\xi) \xi, \xi \rangle| + |\langle f''(x) \xi, \xi \rangle| \right) \mu(\dd\xi) . \end{align*} Here \begin{align*} B^{(1)}_{\theta,M}(x) &\leq x_1 \int_{D_M} \|f''(x+\theta^{-1}\tau\xi) - f''(x)\| \|\xi\|^2 \, \mu(\dd\xi) \\ &\leq \sup_{\xi\in D_M} x_1 \|f''(x+\theta^{-1}\tau\xi) - f''(x)\| \int_{\RR_+\times\RR^d} \|\xi\|^2 \, \mu(\dd\xi) , \end{align*} and note that \ $\Vert A-B\Vert\geq \Vert A\Vert - \Vert B\Vert$, $A,B\in\RR^{p\times p}$, \ yields that \begin{align*} x_1 \|f''(x+\theta^{-1}\tau\xi) - f''(x)\| \leq \|(x_1+\theta^{-1}\tau\xi_1) f''(x+\theta^{-1}\tau\xi) - x_1 f''(x)\| + \theta^{-1}\tau\xi_1 \|f''(x+\theta^{-1}\tau\xi)\| . \end{align*} Further, we have again \[ \{ \tau \xi = \tau(\theta, x, \xi) \xi : \text{$\theta \in \RR_{++}$, \ $x \in \RR_+\times\RR^d$ \ and \ $\xi \in D_M$} \} \subset D_M, \] and hence \begin{align*} \sup_{\xi\in D_M} &\|(x_1+\theta^{-1}\tau\xi_1)f''(x+\theta^{-1}\tau\xi) - x_1f''(x)\|\\ &\leq \sup_{\widetilde\xi\in D_M} \|(x_1+\theta^{-1}\widetilde\xi_1)f''(x+\theta^{-1}\widetilde\xi) - x_1f''(x)\| . \end{align*} Since the function \ $x \mapsto x_1 f''(x)$ \ is uniformly continuous on \ $\RR_+\times\RR^d$, \ there exists a \ $\theta_1\in\RR_{++}$ \ (depending on \ $\varepsilon$ \ and \ $M$) \ such that \[ \sup_{\widetilde\xi\in D_M} \|(x_1+\theta^{-1}\widetilde\xi_1)f''(x+\theta^{-1}\widetilde\xi) - x_1f''(x)\| \int_{\RR_+\times\RR^d} \|\xi\|^2 \, \mu(\dd\xi) < \frac{\varepsilon}{4} \] for all \ $x\in\RR_+\times\RR^d$ \ and \ $\theta \in [\theta_1,\infty)$. \ Moreover, there exists a \ $\theta_2\in\RR_{++}$ \ (depending on \ $\varepsilon$ \ and \ $M$) \ such that \begin{align*} \theta^{-1}\sup_{\xi\in D_M} \tau\xi_1 \|f''(x+\theta^{-1}\tau\xi)\| &\int_{\RR_+\times\RR^d} \|\xi\|^2 \, \mu(\dd\xi)\\ &\leq \theta^{-1} \|f''\|_\infty \sup_{\xi\in D_M} \xi_1 \int_{\RR_+\times\RR^d} \|\xi\|^2 \, \mu(\dd\xi) < \frac{\varepsilon}{4} \end{align*} for all \ $x\in\RR_+\times\RR^d$ \ and \ $\theta \in [\theta_2,\infty)$. \ Consequently, \ $B^{(1)}_{\theta,M}(x) < \frac{\varepsilon}{2}$ \ for all \ $x\in\RR_+\times\RR^d$ \ and \ $\theta \in [\theta_1 + \theta_2, \infty)$. \ Further, \begin{align*} B^{(2)}_{\theta,M}(x) &\leq x_1 \int_{(\RR_+\times\RR^d)\setminus D_M} \left( \|f''(x+\theta^{-1}\tau\xi)\| + \|f''(x)\| \right) \|\xi\|^2 \, \mu(\dd\xi) \\ &\leq \sup_{\xi\in \RR_+\times\RR^d} x_1 \left( \|f''(x+\theta^{-1}\tau\xi)\| + \|f''(x)\| \right) \int_{(\RR_+\times\RR^d)\setminus D_M} \|\xi\|^2 \, \mu(\dd\xi) . \end{align*} Here \[ x_1 \left( \|f''(x+\theta^{-1}\tau\xi)\| + \|f''(x)\| \right) \leq (x_1+\theta^{-1}\tau\xi_1)\|f''(x+\theta^{-1}\tau\xi)\| + x_1\|f''(x)\| , \] hence \[ \sup_{\xi\in \RR_+\times\RR^d} x_1 \left( \|f''(x+\theta^{-1}\tau\xi)\| + \|f''(x)\| \right) \leq 2 \sup_{x\in \RR_+\times\RR^d} x_1\|f''(x)\| . \] Consequently, by \eqref{2help6}, we have \begin{align*} B^{(2)}_{\theta,M}(x) \leq 2 \sup_{x\in \RR_+\times\RR^d} x_1\|f''(x)\| \int_{(\RR_+\times\RR^d)\setminus D_M} \|\xi\|^2 \, \mu(\dd\xi) < \frac{\varepsilon}{2} \end{align*} for all \ $x\in\RR_+\times\RR^d$ \ and \ $\theta\in\RR_{++}$. Putting the pieces together we have \eqref{help4}. Finally, Ethier and Kurtz \cite[Corollary 8.7 on page 232]{EthKur} yields our assertion. Namely, with the notations of part (f) of this corollary (but replacing \ $n$ \ by \ $\theta$), \ let \begin{itemize} \item $G_\theta:= E_\theta:= E:=\RR_+\times\RR^d$ \ for all \ $\theta\in\RR_{++}$, \item $C_a := \cC_c^\infty(\RR_+ \times \RR^d)$ \ which strongly separates points in \ $\RR_+\times\RR^d$ \ (indeed, for every \ $(x_1,x_2)\in \RR_+ \times \RR^d$ \ and \ $\delta\in\RR_{++}$, \ the bump function defined by \ $h_1(u_1,u_2) := \exp\bigl\{- \frac{1}{1-(u_1-x_1)^2} - \frac{1}{1-\|u_2-x_2\|^2} \bigr\}$ \ if \ $|u_1-x_1|<1$ \ and \ $\|u_2-x_2\|<1$ \ with \ $(u_1,u_2)\in \RR_+ \times \RR^d$, \ and \ $h_1(u_1,u_2) := 0$ \ otherwise, satisfies (4.7) on page 113 in Ethier and Kurtz \cite{EthKur}), \item $\eta_\theta : E_\theta \to E$ \ with \ $\eta_\theta(x_1,x_2):=(x_1,x_2)$, \ $(x_1,x_2)\in E_\theta$ \ for all \ $\theta\in\RR_{++}$, \item $\pi_\theta : E \to E_\theta$ \ with \ $\pi_\theta(x_1,x_2):= (x_1,x_2)$, \ $(x_1,x_2)\in E$ \ for all \ $\theta\in\RR_{++}$, \item for each \ $f \in \cC_c^\infty(\RR_+ \times \RR^d)$ \ one can choose \ $f_\theta:=f$ \ and \ $g_\theta := \cA_{(Y^{(\theta)}_\theta, X^{(\theta)}_\theta)} f$ \ for all \ $\theta\in\RR_{++}$ \ (and hence \ $(f_\theta, g_\theta) \in \widehat{\cA}_{(Y^{(\theta)}_\theta,X^{(\theta)}_\theta)}$ \ defined on page 24 in Ethier and Kurtz \cite{EthKur} by part (c) of Proposition 1.5 on page 9 in Ethier and Kurtz \cite{EthKur}), \item $(\cG^\theta_t)_{t\in\RR_+} := (\cF_t^{(Y^{(\theta)}_\theta,X^{(\theta)}_\theta)})_{t\in\RR_+}$, \ where \ $\cF_t^{(Y^{(\theta)}_\theta,X^{(\theta)}_\theta)}$ \ denotes the $\sigma$-algebra generated by \ $\{(Y^{(\theta)}_\theta(s),X^{(\theta)}_\theta(s)), s\in[0,t]\}$. \end{itemize} Then, by our assumptions, convergence of the initial distributions holds, condition (8.35) on page 232 in Ethier and Kurtz \cite{EthKur} is automatically satisfied, and \eqref{help2} shows the validity of condition (8.36) on page 232 in Ethier and Kurtz~\cite{EthKur}. \proofend \begin{Rem} By giving an example, we shed some light on why we consider only \ $(1+d)$-dimensional affine processes with state space \ $\RR_+\times\RR^d$ \ in Theorem \ref{Thm1} instead of \ $(n+d)$-dimensional ones with state space \ $\RR_+^n\times\RR^d$, $n\in\NN$. \ Let \ $(Y_t)_{t\in\RR_+}$ \ be a two-dimensional continuous state branching process with infinitesimal generator \begin{align*} (\cA_Y f)(y) = \sum_{i=1}^2 y_i \int_{\RR_+^2 \setminus\{0\} } \Bigl( f(y+u) - f(y) - f'_i(y) u_i \Bigr) p_i(\dd u), \end{align*} for \ $f\in\cC_c^2(\RR_+^2)$ \ and \ $y=(y_1,y_2)\in\RR_+^2$, \ where \ $p_i$, $i=1,2$, \ are \ $\sigma$-finite measures on \ $\RR_+^2\setminus\{0\}$ \ such that \begin{align}\label{help48} \int_{\RR_+^2\setminus\{0\}} (u_1 + \Vert u\Vert^2) p_2(\dd u) < \infty \qquad \text{and}\qquad \int_{\RR_+^2\setminus\{0\}} (u_2 + \Vert u\Vert^2) p_1(\dd u) < \infty, \end{align} see, e.g., Duffie et al. \cite[Theorem 2.7]{DufFilSch}. Note that \ $Y$ \ can be considered as a two-dimensional affine process with state space \ $\RR_+^2$ \ (formally with \ $d=0$). \ Then, by a simple modification of Lemma \ref{Lem_inf_gen}, for all \ $\theta>0$, \ $f\in\cC_c^2(\RR_+^2)$ \ and \ $y=(y_1,y_2)\in\RR_+^2$, \begin{align*} (\cA_{Y_\theta}f)(y) &= \theta(\cA_Y f_\theta)(\theta y)\\ &\quad= \theta \sum_{i=1}^2 \theta y_i \int_{\RR_+^2\setminus\{0\}} \Bigl( f(\theta^{-1}(\theta y +u)) - f(\theta^{-1}\theta y) - \theta^{-1} f'_i(\theta^{-1} \theta y) u_i \Bigr) p_i(\dd u) \\ &\quad= \theta^2 \sum_{i=1}^2 y_i \int_{\RR_+^2\setminus\{0\}} \Bigl( f(y + \theta^{-1}u) - f(y) - \langle f'(y), \theta^{-1}u\rangle \Bigr) p_i(\dd u)\\ &\phantom{\quad=\;} + \theta \sum_{i=1}^2 y_i f'_{3-i}(y) \int_{\RR_+^2\setminus\{0\}} u_{3-i} \, p_i(\dd u), \end{align*} where the last equality follows by \eqref{help48}. \ Supposing that \ $f$ \ is real-valued, by Taylor's theorem, \begin{align*} f(y + \theta^{-1}u) - f(y) - \langle f'(y), \theta^{-1}u\rangle = \frac{1}{2}\langle f''(y+\tau\theta^{-1}u) \theta^{-1} u, \theta^{-1}u\rangle = \frac{\theta^{-2}}{2}\langle f''(y+\tau\theta^{-1}u) u, u\rangle \end{align*} with some \ $\tau=\tau(u,y)\in[0,1]$. \ Hence, similarly to the proof of \eqref{help4}, we get \begin{align*} \lim_{\theta\to\infty} \theta^2 &\sum_{i=1}^2 y_i \int_{\RR_+^2\setminus\{0\}} \Bigl( f(y + \theta^{-1}u) - f(y) - \langle f'(y), \theta^{-1}u\rangle \Bigr) p_i(\dd u)\\ & = \frac{1}{2} \sum_{i=1}^2 y_i \int_{\RR_+^2\setminus\{0\}} \langle f''(y)u,u\rangle p_i(\dd u) \end{align*} for real-valued \ $f\in\cC_c^2(\RR_+^2)$ \ and \ $y=(y_1,y_2)\in\RR_+^2$. \ However, \ $(\cA_{Y_\theta}f)(y)$ \ does not converge as \ $\theta \to \infty$ \ provided that \[ \sum_{i=1}^2 y_i f'_{3-i}(y) \int_{\RR_+^2\setminus\{0\}} u_{3-i} \, p_i(\dd u) \ne 0. \] We also note that this phenomena is somewhat similar to that of Remark 2.1 in Ma \cite{Ma}.\proofend \end{Rem} In the next remark we formulate some special cases of Theorem \ref{Thm1}. \begin{Rem}\label{Rem1} (i) If \ $(Y(t),X(t))_{t\in\RR_+}$ \ is a \ $(1+d)$-dimensional affine process on \ $\RR_+\times\RR^d$ \ with admissible parameters \ $(a,\alpha,b,0,m,\mu)$ \ such that condition \eqref{cond1} is satisfied, then the conditions of Theorem \ref{Thm1} are satisfied for \ $(Y^{(\theta)}(t),X^{(\theta)}(t))_{t\in\RR_+} := (Y(t),X(t))_{t\in\RR_+}$, \ $\theta\in\RR_{++}$, \ and hence \begin{align*} \left( \theta^{-1} Y(\theta t), \theta^{-1} X(\theta t)\right)_{t\in\RR_+} \distr (\cY(t),\cX(t))_{t\in\RR_+} \qquad \text{as \ $\theta\to\infty$} \end{align*} in \ $\DD(\RR_+,\RR_+\times\RR^d)$, \ where \ $(\cY(t),\cX(t))_{t\in\RR_+}$ \ is a \ $(1+d)$-dimensional affine process on \ $\RR_+\times\RR^d$ \ with admissible parameters \ $(0,\widetilde{\alpha},\widetilde{b},0,0,0)$, \ where \ $\widetilde{\alpha}$ \ and \ $\widetilde{b}$ \ are given in Theorem \ref{Thm1}. (ii) If \ $(Y(t),X(t))_{t\in\RR_+}$ \ is a \ $(1+d)$-dimensional affine process on \ $\RR_+\times\RR^d$ \ with \ $(Y(0),X(0))=(0,0)$ \ and with admissible parameters \ $(0,\alpha,b,0,0,0)$, \ then \[ \left(\theta^{-1} Y(\theta t), \theta^{-1} X(\theta t) \right)_{t\in\RR_+} \distre (Y(t),X(t))_{t\in\RR_+} \qquad \text{for all \ $\theta\in\RR_{++}$,} \] where \ $\distre$ \ denotes equality in distribution. Indeed, by Proposition 1.6 on page 161 in Ethier and Kurtz \cite{EthKur}, it is enough to check that the semigroups (on the Banach space of bounded Borel measurable functions on \ $\RR_+\times\RR^d$) \ corresponding to the processes in question coincide. By the definition of a core, this follows from the equality of the infinitesimal generators of the processes in question on the core \ $\cC_c^\infty(\RR_+\times\RR^d)$, \ which has been shown in the proof of Theorem \ref{Thm1}.\proofend \end{Rem} Next we present a corollary of Theorem \ref{Thm1} which states weak convergence of appropriately normalized one-dimensional continuous state branching processes with immigration. Our corollary generalizes Theorem 2.3 in Huang et al.\ \cite{HuaMaZhu} in the sense that we do not have to suppose that \ $\int_1^\infty \xi^2\,m(\dd\xi)<\infty$, \ only that \ $\int_1^\infty \xi \,m(\dd\xi)<\infty$ \ (with the notations of Huang et al.\ \cite{HuaMaZhu}), and our proof defers from that of Huang et al. \cite{HuaMaZhu}. \begin{Cor} For all \ $\theta\in\RR_{++}$, \ let \ $(Y^{(\theta)}(t))_{t\in\RR_+}$ \ be a one-dimensional continuous state branching process with immigration on \ $\RR_+$ \ with branching mechanism \[ R^{(\theta)}(z) := \beta^{(\theta)}z + \alpha^{(\theta)}z^2 + \int_{\RR_+} (\ee^{-zu} - 1 + zu)\,p(\dd u), \qquad z\in\RR_+, \] and with immigration mechanism \[ F^{(\theta)}(z):= b^{(\theta)}z + \int_{\RR_+} (1- \ee^{-zu} )\,n(\dd u), \qquad z\in\RR_+, \] where \ $\alpha^{(\theta)}\geq 0$, \ $b^{(\theta)}\geq 0$, \ $\beta^{(\theta)}\in\RR$ \ and \ $n$ \ and \ $p$ \ are measures on \ $(0,\infty)$ \ such that \[ \int_{\RR_+} u \, n(\dd u)<\infty \qquad \text{and}\qquad \int_{\RR_+} u^2 \, p(\dd u)<\infty. \] Let \ $\alpha, b, \beta \in \RR$, \ and let \ $(Y(t))_{t\in\RR_+}$ \ be a one-dimensional continuous state branching process with immigration on \ $\RR_+$ \ with branching mechanism \[ R(z):= -\beta z + \left(\alpha + \frac{1}{2}\int_{\RR_+} u^2\, p(\dd u)\right)z^2, \qquad z\in\RR_+, \] and with immigration mechanism \[ F(z):= \left(b + \int_{\RR_+} u\, n(\dd u) \right)z,\qquad z\in\RR_+. \] If \[ \lim_{\theta\to\infty} \alpha^{(\theta)} = \alpha, \qquad \lim_{\theta\to\infty} b^{(\theta)} = b, \qquad \lim_{\theta\to\infty}\theta \beta^{(\theta)}=\beta, \qquad Y^{(\theta)}(0)\distr Y(0) \] as \ $\theta\to\infty$, \ then \begin{align*} \left(\theta^{-1} Y^{(\theta)}(\theta t)\right)_{t\in\RR_+} \distr (Y(t))_{t\in\RR_+} \qquad \text{as \ $\theta\to\infty$} \end{align*} in \ $\DD(\RR_+,\RR_+)$. \end{Cor} \noindent{\bf Proof.} For each \ $\theta \in \RR_{++}$ \ and \ $t \in \RR_+$, \ let \ $X^{(\theta)}(t) := 0$. \ Then for each \ $\theta \in \RR_{++}$, \ the process \ $(Y^{(\theta)}(t),X^{(\theta)}(t))_{t\in\RR_+}$ \ is a two-dimensional affine process with admissible parameters \ $(0,\overline{\alpha}^{(\theta)},\overline{b}^{(\theta)}, \overline{\beta}^{(\theta)},m,\mu)$, \ where \begin{gather*} \overline{\alpha}^{(\theta)} := \begin{bmatrix} \alpha^{(\theta)} & 0 \\ 0 & 0 \\ \end{bmatrix} , \qquad \overline{b}^{(\theta)} := \begin{bmatrix} b^{(\theta)} \\ 0 \end{bmatrix} , \qquad \overline{\beta}^{(\theta)} := \begin{bmatrix} \beta^{(\theta)} & 0 \\ 0 & 0 \\ \end{bmatrix} , \\ \mu(\dd\xi) = \mu(\dd\xi_1,\dd\xi_2) := p(\dd\xi_1)\times \delta_0(\dd\xi_2),\\ m(\dd\xi) = m(\dd\xi_1,\dd\xi_2) := n(\dd\xi_1)\times \delta_0(\dd\xi_2), \end{gather*} where \ $\delta_0$ \ denotes the Dirac measure concentrated on \ $0\in\RR$. \ Then, by Theorem \ref{Thm1}, for the two-dimensional affine processes \ $(Y^{(\theta)}(t),X^{(\theta)}(t))_{t\in\RR_+}$, \ $\theta \in \RR_{++}$, \ we have \begin{align*} \left( \theta^{-1} Y^{(\theta)}(\theta t), \theta^{-1} X^{(\theta)}(\theta t) \right)_{t\in\RR_+} \distr (Y(t),X(t))_{t\in\RR_+} \qquad \text{as \ $\theta\to\infty$} \end{align*} in \ $\DD(\RR_+,\RR_+\times\RR)$, \ where \ $(Y(t),X(t))_{t\in\RR_+}$ \ is a two-dimensional affine process on \ $\RR_+\times\RR$ \ with infinitesimal generator \begin{align*} (\cA_{(Y,X)} f)(x) = \left(\alpha + \frac{1}{2}\int_{\RR_+}u^2\,p(\dd u)\right)x_1f_{1,1}''(x) + \left(b + \beta x_1 + \int_{\RR_+}u\,n(\dd u)\right)f_1'(x), \end{align*} for \ $x=(x_1,x_2)\in\RR_+\times\RR$ \ and \ $f\in\cC^2_c(\RR_+\times\RR)$. \ Note that in fact \ $X$ \ is the identically zero process. Finally, Theorem 9.30 in Li \cite{Li} yields the assertion. \proofend \section{Least squares estimator for a critical two-dimensional affine diffusion process}\label{Section_statistics} In this section continuous time stochastic processes will be written as \ $(\xi_t)_{t\in\RR_+}$ \ instead of \ $(\xi(t))_{t\in\RR_+}$. \ Let \ $(\Omega,\cF,(\cF_t)_{t\in\RR_+},\PP)$ \ be a filtered probability space satisfying the usual conditions, i.e., \ $(\Omega,\cF,\PP)$ \ is complete, the filtration $(\cF_t)_{t\in\RR_+}$ \ is right-continuous and \ $\cF_0$ \ contains all the \ $\PP$-null sets in \ $\cF$. \ Let \ $(W_t)_{t\in\RR_+}$ \ and \ $(B_t)_{t\in\RR_+}$ \ be independent standard \ $(\cF_t)_{t\in\RR_+}$-Wiener processes. Let us consider the following two-dimensional diffusion process given by the SDE \begin{align}\label{2dim_affine} \begin{cases} \dd Y_t = (a-bY_t)\,\dd t + \sqrt{Y_t}\,\dd W_t,\\ \dd X_t = (m-\theta X_t)\,\dd t + \sqrt{Y_t}\,\dd B_t, \end{cases} \qquad t\in\RR_+, \end{align} where \ $a\in\RR_{++}$ \ and \ $b,\theta,m\in\RR$. \subsection{Preparations and (sub)(super)criticality} The next proposition is about the existence and uniqueness of a strong solution of the SDE \eqref{2dim_affine}. \begin{Pro}\label{Pro_affine} Let \ $(\eta,\zeta)$ \ be a random vector independent of \ $(W_t,B_t)_{t\in\RR_+}$ \ satisfying \ $\PP(\eta\geq 0)=1$. \ Then, for all \ $a\in\RR_{++}$ \ and \ $b, m, \theta\in\RR$, \ there is a (pathwise) unique strong solution \ $(Y_t,X_t)_{t\in\RR_+}$ \ of the SDE \eqref{2dim_affine} such that \ $\PP((Y_0,X_0) = (\eta,\zeta))=1$ \ and \ $\PP(\text{$Y_t\geq 0$ \ for all \ $t\in\RR_+$})=1$. \ Further, for all \ $0\leq s\leq t$, \begin{align}\label{help27} Y_t = \ee^{-b(t-s)} \left( Y_s + a\int_s^t \ee^{-b(s-u)}\,\dd u + \int_s^t \ee^{-b(s-u)}\sqrt{Y_u} \,\dd W_u\right), \end{align} and \begin{align}\label{help22} X_t = \ee^{-\theta(t-s)} \left( X_s + m\int_s^t \ee^{-\theta(s-u)}\,\dd u + \int_s^t \ee^{-\theta(s-u)}\sqrt{Y_u} \,\dd B_u\right). \end{align} \end{Pro} \noindent{\bf Proof.} By Ikeda and Watanabe \cite[Example 8.2, page 221]{IkeWat}, there is a pathwise unique non-negative strong solution \ $(Y_t)_{t\in\RR_+}$ \ of the first equation in \eqref{2dim_affine} with any initial value \ $\eta$ \ satisfying \ $\PP(\eta\geq 0)=1$. \ Next, by applications of the It\^{o}'s formula to the processes \ $(\ee^{b t} Y_t)_{t\in\RR_+}$ \ and \ $(\ee^{\theta t} X_t)_{t\in\RR_+}$, \ respectively, we have \begin{align*} \dd(\ee^{b t} Y_t) &= b\ee^{b t}Y_t\,\dd t + \ee^{b t}\dd Y_t = b \ee^{b t}Y_t\,\dd t + \ee^{b t}\big( (a-b Y_t)\,\dd t + \sqrt{Y_t}\,\dd W_t\big)\\ &= a\ee^{b t}\,\dd t + \ee^{b t}\sqrt{Y_t}\,\dd W_t, \qquad t\in\RR_+, \end{align*} and \begin{align*} \dd(\ee^{\theta t} X_t) &= \theta\ee^{\theta t}X_t\,\dd t + \ee^{\theta t}\dd X_t = \theta\ee^{\theta t}X_t\,\dd t + \ee^{\theta t}\big( (m-\theta X_t)\,\dd t + \sqrt{Y_t}\,\dd B_t\big)\\ &= m\ee^{\theta t}\,\dd t + \ee^{\theta t}\sqrt{Y_t}\,\dd B_t, \qquad t\in\RR_+, \end{align*} which imply \eqref{help27} and \eqref{help22} in case of \ $s=0$. \ If \ $0\leq s\leq t$, \ then, by \[ \ee^{b t} Y_t = \ee^{b s} Y_s + a\int_s^t \ee^{b u}\,\dd u + \int_s^t \ee^{b u}\sqrt{Y_u} \,\dd B_u, \] and \[ \ee^{\theta t} X_t = \ee^{\theta s} X_s + m\int_s^t \ee^{\theta u}\,\dd u + \int_s^t \ee^{\theta u}\sqrt{Y_u} \,\dd B_u, \] we have \eqref{help27} and \eqref{help22}. Finally, we note that the existence of a pathwise unique strong solution \ $(Y_t,X_t)_{t\in\RR_+}$ \ of the SDE \eqref{2dim_affine} with \ $\PP(\text{$Y_t\geq 0$ \ for all \ $t\in\RR_+$})=1$ \ follows also by a general result of Dawson and Li \cite[Theorem 6.2]{DawLi}. \proofend Note that it is the assumption \ $a\in\RR_{++}$ \ that ensures \ $\PP(\text{$Y_t\,{\geq}\, 0$, \ $\forall\,t\,{\in}\,\RR_+$})\,{=}\,1$. Next we present a result about the first moment of \ $(Y_t,X_t)_{t\in\RR_+}$. \begin{Pro}\label{Pro_moments} Let \ $(Y_t,X_t)_{t\in\RR_+}$ \ be a strong solution of the SDE \eqref{2dim_affine} satisfying \ $\PP(Y_0\geq 0)=1$, \ $\EE(Y_0)<\infty$, \ and \ $\EE(X_0)<\infty$. \ Then \begin{align*} \begin{bmatrix} \EE(Y_t) \\ \EE(X_t) \\ \end{bmatrix} = \begin{bmatrix} \ee^{-bt} & 0 \\ 0 & \ee^{-\theta t} \\ \end{bmatrix} \begin{bmatrix} \EE(Y_0) \\ \EE(X_0) \\ \end{bmatrix} + \begin{bmatrix} \int_0^t \ee^{-bs} \, \dd s & 0 \\ 0 & \int_0^t \ee^{-\theta s} \, \dd s \\ \end{bmatrix} \begin{bmatrix} a \\ m \\ \end{bmatrix},\qquad t\in\RR_+, \end{align*} \end{Pro} \noindent{\bf Proof.} By Proposition \ref{Pro_affine}, we have \begin{align*} &Y_t = \ee^{-bt} \left( Y_0 + a\int_0^t\ee^{bu}\,\dd u + \int_0^t \ee^{bu} \sqrt{Y_u}\,\dd W_u\right), \qquad t\in\RR_+,\\ &X_t = \ee^{-\theta t} \left( X_0 + m\int_0^t\ee^{\theta u}\,\dd u + \int_0^t \ee^{\theta u} \sqrt{Y_u}\,\dd B_u\right), \qquad t\in\RR_+, \end{align*} and so, taking expectations of both sides, \begin{align*} & \EE(Y_t) = \ee^{-bt}\EE(Y_0) + a \ee^{-bt} \int_0^t\ee^{bu}\,\dd u = \ee^{-bt}\EE(Y_0) + a \int_0^t\ee^{-bu}\,\dd u, \quad t\in\RR_+,\\ & \EE(X_t) = \ee^{-\theta t}\EE(X_0) + m \ee^{-\theta t} \int_0^t\ee^{\theta u}\,\dd u = \ee^{-\theta t}\EE(X_0) + m \int_0^t\ee^{-\theta u}\,\dd u, \;\; t\in\RR_+, \end{align*} where we used that the processes \[ \left( \int_0^t \ee^{bu} \sqrt{Y_u}\,\dd W_u \right)_{t\in\RR_+} \quad \text{and}\qquad \left( \int_0^t \ee^{\theta u} \sqrt{Y_u}\,\dd B_u \right)_{t\in\RR_+} \] are martingales which can be checked as follows. First we check that they are local martingales with respect to the filtration \ $(\cF_t)_{t\in\RR_+}$. \ Let us define the increasing sequence of stopping times by \ $\delta_n:=\inf\{t\geq 0:Y_t\geq n\}$, $n\in\NN$. \ Since \ $Y$ \ has continuous trajectories almost surely, we have \ $\PP(\lim_{n\to\infty}\delta_n=\infty)=1$. \ Using \ $(\delta_n)_{n\in\NN}$ \ as a localizing sequence, we have \[ \EE\left(\int_0^{t\wedge \delta_n} \ee^{2bu} Y_u\,\dd u \right) \leq nt\max(1,\ee^{2bt}), \quad t\in\RR_+,\; n\in\NN. \] The local martingale property of \ $\bigl( \int_0^t \ee^{bu} \sqrt{Y_u}\,\dd W_u \bigr)_{t\in\RR_+}$ \ follows by Ikeda and Watanabe \cite[page 57]{IkeWat}. Hence, using \eqref{help27} and that \ $a\in\RR_{++}$, \ we find that \[ \EE(\ee^{b(t\wedge \delta_n)}Y_{t\wedge \delta_n}) = \EE(Y_0) + a \EE\left( \int_0^{t\wedge \delta_n} \ee^{bu}\,\dd u \right) \leq \EE(Y_0) + at\max(1,\ee^{bt}) \] for all \ $t\in\RR_+$ \ and \ $n\in\NN$, \ and then, by Fatou's lemma, \begin{align}\label{help42} \EE(\ee^{bt}Y_t) \leq \liminf_{n\to\infty} \EE(\ee^{b(t\wedge \delta_n)}Y_{t\wedge \delta_n}) \leq \EE(Y_0) + at\max(1,\ee^{bt}), \qquad t\in\RR_+. \end{align} Next, we can deduce that \ $\bigl( \int_0^t \ee^{bu} \sqrt{Y_u}\,\dd W_u \bigr)_{t\in\RR_+}$ \ is indeed a martingale.\break First, we note that a local martingale \ $M$ \ is a square integrable martingale if \ $E([M,M]_t)<\infty$ \ for all \ $t\in\RR_+$, \ where \ $([M,M]_t)_{t\in\RR_+}$ \ denotes the quadratic variation process of \ $M$, \ see, e.g., Corollary 3 on page 73 in Protter \cite{Pro}. Here the quadratic variation process of \ $\bigl( \int_0^t \ee^{bu} \sqrt{Y_u}\,\dd W_u \bigr)_{t\in\RR_+}$ \ takes the form \[ E\left(\int_0^t \ee^{2bu} Y_u\,\dd u \right) <\infty, \qquad t\in\RR_{+}, \] where, for the inequality, we used Fubini's theorem, \eqref{help42} and our assumption \ $\EE(Y_0)<\infty$. Replacing \ $b$ \ by \ $\theta$, \ we have the desired martingale property of \ $\bigl( \int_0^t \ee^{\theta u} \sqrt{Y_u}\,\dd B_u \bigr)_{t\in\RR_+}$, \ too. \proofend Next we show that the process \ $(Y_t,X_t)_{t\in\RR_+}$ \ given by the SDE \eqref{2dim_affine} is an affine process. \begin{Pro}\label{Pro_affine_2} Let \ $(Y_t,X_t)_{t\in\RR_+}$ \ be a strong solution of the SDE \eqref{2dim_affine} satisfying \ $\PP(Y_0\geq 0)=1$. \ Then \ $(Y_t,X_t)_{t\in\RR_+}$ \ is an affine process with infinitesimal generator \begin{align}\label{help50_affine_generator} \begin{split} (\cA_{(Y,X)} f)(x) = (a-bx_1)f_1'(x) + (m-\theta x_2)f_2'(x) + \frac{1}{2}x_1(f_{1,1}''(x) + f_{2,2}''(x)) \end{split} \end{align} for \ $x=(x_1,x_2)\in\RR_+\times\RR$ \ and \ $f\in\cC^2_c(\RR_+\times\RR)$. \end{Pro} \noindent{\bf Proof.} For calculating the infinitesimal generator of \ $(Y_t,X_t)_{t\in\RR_+}$, \ without loss of generality, we may suppose that \ $\PP((Y_0,X_0) = (y_0,x_0))=1$, \ where \ $(y_0,x_0)\in\RR_+\times\RR$. \ By It\^{o}'s formula, for all real-valued functions \ $f\in\cC^2_c(\RR_+\times\RR)$ \ we have \begin{align*} f(Y_t,X_t) & = f(y_0,x_0) + \int_0^t f_1'(Y_s,X_s)\sqrt{Y_s} \,\dd W_s + \int_0^t f_2'(Y_s,X_s)\sqrt{Y_s} \,\dd B_s \\ &\phantom{=\;} + \int_0^t f_1'(Y_s,X_s)(a-bY_s)\,\dd s + \int_0^t f_2'(Y_s,X_s)(m-\theta X_s)\,\dd s \\ &\phantom{=\;} + \frac{1}{2}\left(\int_0^t f_{1,1}''(Y_s,X_s)Y_s\,\dd s + \int_0^t f_{2,2}''(Y_s,X_s)Y_s\,\dd s \right) \\ &= f(y_0,x_0) + \int_0^t (\cA_{(Y,X)} f)(Y_s,X_s)\,\dd s + M_t(f),\qquad t\in\RR_+, \end{align*} where \[ M_t(f):= \int_0^t f_1'(Y_s,X_s)\sqrt{Y_s} \,\dd W_s + \int_0^t f_2'(Y_s,X_s)\sqrt{Y_s} \,\dd B_s,\qquad t\in\RR_+, \] and \ $\cA_{(Y,X)} f$ \ is given by \eqref{help50_affine_generator}. Here \ $(M_t(f))_{t\in\RR_+}$ \ is a square integrable martingale with respect to the filtration \ $(\cF_t)_{t\in\RR_+}$, \ since \begin{align*} &\int_0^t \EE((f_1'(Y_s,X_s))^2Y_s) \,\dd s \leq C_1 \int_0^t \EE(Y_s) \,\dd s <\infty,\qquad t\in\RR_+,\\ & \int_0^t \EE((f_2'(Y_s,X_s))^2Y_s) \,\dd s \leq C_2 \int_0^t \EE(Y_s) \,\dd s <\infty,\qquad t\in\RR_+, \end{align*} with some constants \ $C_1>0$ \ and \ $C_2>0$, \ where the finiteness of the integrals follow by Proposition \ref{Pro_moments}. Finally, if \ $f\in\cC^2_c(\RR_+\times\RR)$ \ is complex valued, then, by decomposing \ $f$ \ into real and imaginary parts, one can argue in the same way as above. \proofend By Proposition \ref{Pro_affine_2}, the process \ $(Y_t,X_t)_{t\in\RR_+}$ \ given by \eqref{2dim_affine} is a two-dimen\-sional affine process with admissible parameters \[ \left( \begin{bmatrix} 0 & 0 \\ 0 & 0 \end{bmatrix} , \, \frac{1}{2} \begin{bmatrix} 1 & 0 \\ 0 & 1 \end{bmatrix} , \, \begin{bmatrix} a \\ m \end{bmatrix} , \, \begin{bmatrix} -b & 0 \\ 0 & -\theta \\ \end{bmatrix} , \, 0 , 0 \right) . \] In what follows we define and study criticality of the affine process given by the SDE \eqref{2dim_affine}. \begin{Def}\label{Def_criticality} Let \ $(Y_t,X_t)_{t\in\RR_+}$ \ be an affine diffusion process given by the SDE \eqref{2dim_affine} satisfying \ $\PP(Y_0\geq 0)=1$. \ We call \ $(Y_t,X_t)_{t\in\RR_+}$ \ subcritical, critical or supercritical if the spectral radius of the matrix \[ \begin{pmatrix} \ee^{-bt} & 0 \\ 0 & \ee^{-\theta t} \\ \end{pmatrix} \] is less than \ $1$, \ equal to \ $1$ \ or greater than \ $1$, \ respectively. \end{Def} Note that, since the spectral radius of the matrix given in Definition \ref{Def_criticality} is \ $\max(\ee^{-bt},\ee^{-\theta t})$, \ the affine process given in Definition \ref{Def_criticality} is \begin{align*} \text{subcritical} \qquad & \text {if \ $b>0$ \ and \ $\theta>0$,}\\ \text{critical} \qquad & \text{if \ $b=0$, \ $\theta\geq 0$ \ or \ $b\geq 0$, \ $\theta=0$,}\\ \text{supercritical} \qquad & \text{if \ $b<0$ \ or \ $\theta<0$.} \end{align*} Definition \ref{Def_criticality} of criticality is in accordance with the corresponding definition for one-dimensional continuous state branching processes, see, e.g., Li \cite[page~58]{Li}. In this section we will always suppose that \begin{align*} \textbf{Condition (C): } &\quad (b,\theta)=(0,0),\; \PP(Y_0 \geq 0)=1,\\ &\quad \text{$\EE(Y_0) < \infty$, \ and \ $\EE(X_0^2) < \infty$.} \end{align*} For some explanations why we study only this special case, see Remarks \ref{Rem2}, \ref{Rem3} and \ref{Rem6}. In the next sections under Condition (C) we will study asymptotic behaviour of least squares estimator of \ $\theta$ \ and \ $(\theta,m)$, \ respectively. Before doing so we recall some critical models both in discrete and continuous time. In general, parameter estimation for critical models has a long history. A common feature of the estimators for parameters of critical models is that one may prove weak limit theorems for them by using norming factors that are usually different from the norming factors for the subcritical and supercritical models. Further, it may happen that one has to use different norming factors for two different critical cases. We recall some discrete time critical models. If \ $(\xi_k)_{k\in\ZZ_+}$ \ is an AR(1) process, i.e., \ $\xi_k = \varrho \xi_{k-1} +\zeta_k$, \ $k\in\NN$, \ with \ $\xi_0 = 0$ \ and an i.i.d.\ sequence \ $(\zeta_k)_{k\in\NN}$ \ having mean \ $0$ \ and positive variance, then the (ordinary) least squares estimator of the so-called stability parameter \ $\varrho$ \ based on the sample \ $\xi_1,\ldots,\xi_n$ \ takes the form \[ \widetilde\varrho_n = \frac{\sum_{k=1}^n \xi_{k-1}\xi_k}{\sum_{k=1}^n \xi_k^2}, \qquad n\in\NN, \] see, e.g., Hamilton \cite[17.4.2]{Ham}. In the critical case, i.e., when \ $\varrho=1$, \ by Hamilton \cite[17.4.7]{Ham}, \[ n(\widetilde\varrho_n-1) \distr \frac{\int_0^1 \cW_t \, \dd \cW_t}{\int_0^1 \cW_t^2 \, \dd t} \qquad \text{as \ $n\to\infty$,} \] where \ $(\cW_t)_{t\in\RR_+}$ \ is a standard Wiener process and \ $\distr$ \ denotes convergence in distribution. Here \ $n(\widetilde\varrho_n-1)$ \ is known as the Dickey-Fuller statistics. We emphasize that the asymptotic behaviour of \ $\widetilde\varrho_n$ \ is completely different in the subcritical \ $(\vert\rho\vert<1)$ \ and supercritical \ $(\vert\rho\vert>1)$ \ cases, where it is asymptotically normal and asymptotically Cauchy, respectively, see, e.g., Mann and Wald \cite{ManWal}, Anderson \cite{And} and White \cite{Whi}. For continuous time critical models, we recall that Huang et al.\ \cite[Theorem 2.4]{HuaMaZhu} studied asymptotic behaviour of weighted conditional least squares estimator of the drift parameters for discretely observed continuous time critical branching processes with immigration given by \begin{align*} \widetilde Y_t & = \widetilde Y_0 + \int_0^t(a+b \widetilde Y_s)\,\dd s + \sigma \int_0^t \sqrt{\widetilde Y_s}\,\dd \cW_s + \int_0^t \int_{[0,\infty)}\xi\,\cN_0(\dd s,\dd \xi)\\ &\phantom{=\;} + \int_0^t \int_0^{\widetilde Y_{s-}}\int_{[0,\infty)}\xi\,(\cN_1(\dd s,\dd u,\dd \xi) - \dd s\,\dd u\,p(\dd\xi)), \qquad t\in\RR_+, \end{align*} where \ $\widetilde Y_0\geq 0$, \ $a\geq 0$, \ $b\in\RR$, $\sigma\geq 0$, \ $\cW$ \ is a standard Wiener process, \ $\cN_0(\dd s,\dd\xi)$ \ is a Poisson random measure on \ $(0,\infty)\times[0,\infty)$ \ with intensity \ $\dd s\,n(\dd\xi)$, \ $\cN_1(\dd s,\dd u,\dd\xi)$ \ is a Poisson random measure on \ $(0,\infty)\times (0,\infty) \times [0,\infty)$ \ with intensity \ $\dd s\,\dd u\,p(\dd\xi)$ \ such that the \ $\sigma$-finite measures \ $n$ \ and \ $p$ \ are supported by \ $(0,\infty)$ \ and \[ \int_0^\infty \xi\, n(\dd\xi) + \int_0^\infty \xi\wedge\xi^2\,p(\dd\xi) <\infty. \] Our technique differs from that of Huang et al.\ \cite{HuaMaZhu} and for completeness we note that the limit distribution and some parts of the proof of their Theorem 2.4 suffer from some misprints. Furthermore, Hu and Long \cite{HuLon3} studied the problem of parameter estimation for critical mean-reverting $\alpha$-stable motions \[ \dd \widetilde X_t = (m-\theta \widetilde X_t)\,\dd t + \dd Z_t,\qquad t\in\RR_+, \] where \ $Z$ \ is an \ $\alpha$-stable L\'evy motion with \ $\alpha\in(0,2)$) \ observed at discrete instants. A least squares estimator is obtained and its asymptotics is discussed in the singular case \ $(m,\theta)=(0,0)$. \ We note that the forms of the limit distributions of least squares estimators for critical two-dimensional affine diffusion processes in our Theorems \ref{Thm2} and \ref{Thm3} are the same as that of the limit distributions in Theorems 3.2 and 4.1 in Hu and Long \cite{HuLon3}, respectively. We also recall that Hu and Long \cite{HuLon1} considered the problem of parameter estimation not only for critical mean-reverting $\alpha$-stable motions, but also for some subcritical ones \ ($m=0$ \ and \ $\theta>0$) \ by proving limit theorems for the least squares estimators that are completely different from the ones in the critical case. \ Huang et al.\ \cite{HuaMaZhu} investigated the asymptotic behaviour of weighted conditional least squares estimator of the drift parameters not only for critical continuous time branching processes with immigration, but also for subcritical and supercritical ones. Using our scaling Theorem \ref{Thm1} we can only handle a special critical affine diffusion model given by \eqref{DD} (for a more detailed discussion, see Remark \ref{Rem3}). The other critical and non-critical cases are under investigation but different techniques are needed. \subsection{Least squares estimator of \ $\theta$ \ when \ $m$ \ is known} \label{section_LSE_theta} The least squares estimator (LSE) of \ $\theta$ \ based on the observations \ $X_i$, $i=0,1,\ldots,n$, \ can be obtained by solving the extremum problem \[ \widetilde\theta_n^{\mathrm{LSE}} := \argmin_{\theta\in\RR} \sum_{i=1}^n (X_i - X_{i-1} - (m-\theta X_{i-1}))^2. \] This definition of LSE of \ $\theta$ \ can be considered as the counterpart of the one given in Hu and Long \cite[formula (1.2)]{HuLon1} for generalized Ornstein-Uhlenbeck processes driven by \ $\alpha$-stable motions, see also Hu and Long \cite[formulas (3.1) and (4.1)]{HuLon3}. For a mathematical motivation of the definition of the LSE of \ $\theta$, \ see later on Remark \ref{Rem5}. With the notation \ $f(\theta):=\sum_{i=1}^n (X_i - X_{i-1} - (m-\theta X_{i-1}))^2$, \ $\theta\in\RR$, \ the equation \ $f'(\theta)=0$ \ takes the form: \[ 2 \sum_{i=1}^n (X_i - X_{i-1} - (m-\theta X_{i-1})) X_{i-1} = 0. \] Hence \[ \left(\sum_{i=1}^n X_{i-1}^2\right)\theta = - \sum_{i=1}^n (X_i - X_{i-1} - m) X_{i-1}, \] i.e., \begin{align}\label{help11} \begin{split} \widetilde\theta_n^{\mathrm{LSE}} = - \frac{\sum_{i=1}^n (X_i - X_{i-1} - m) X_{i-1}}{\sum_{i=1}^n X_{i-1}^2} = - \frac{\sum_{i=1}^n (X_i - X_{i-1})X_{i-1} - \left(\sum_{i=1}^n X_{i-1}\right)m} {\sum_{i=1}^n X_{i-1}^2} \end{split} \end{align} provided that \ $\sum_{i=1}^n X_{i-1}^2>0$. \ Since \ $f''(\theta)=2\sum_{i=1}^n X_{i-1}^2$, \ $\theta\in\RR$, \ we have \ $\widetilde\theta_n^{\mathrm{LSE}}$ \ is indeed the solution of the extremum problem provided that \ $\sum_{i=1}^n X_{i-1}^2>0$. \begin{Thm}\label{Thm2} Let us assume that Condition (C) holds. Then \ $\PP(\sum_{i=1}^n X_{i-1}^2>0)=1$ \ for all \ $n\geq2$, \ and there exists a unique LSE \ $\widetilde\theta_n^{\mathrm{LSE}}$ \ which has the form given in \eqref{help11}. Further, \begin{align}\label{help23} n\widetilde\theta_n^{\mathrm{LSE}} \distr - \frac{\int_0^1 \cX_t\,\dd \cX_t - m\int_0^1 \cX_t\,\dd t} {\int_0^1 \cX_t^2\,\dd t} \qquad \text{as \ $n\to\infty$,} \end{align} where \ $(\cX_t)_{t\in\RR_+}$ \ is the second coordinate of a two-dimensional affine process \ $(\cY_t,\cX_t)_{t\in\RR_+}$ \ given by the unique strong solution of the SDE \begin{align}\label{help16} \begin{split} \begin{cases} \dd\cY_t = a\,\dd t + \sqrt{\cY_t}\,\dd \cW_t,\\[2mm] \dd\cX_t = m\,\dd t + \sqrt{\cY_t}\,\dd \cB_t, \end{cases}\qquad t\in\RR_+, \end{split} \end{align} with initial value \ $(\cY_0,\cX_0)=(0,0)$, \ where \ $(\cW_t)_{t\in\RR_+}$ \ and \ $(\cB_t)_{t\in\RR_+}$ \ are independent standard Wiener processes. \end{Thm} \begin{Rem}\label{Rem2} (i) The limit distributions in Theorem \ref{Thm2} have the same forms as those of the limit distributions in Theorem 3.2 in Hu and Long \cite{HuLon3}. \noindent (ii) The limit distribution of \ $n\widetilde\theta_n^{\mathrm{LSE}}$ \ as \ $n\to\infty$ \ in Theorem \ref{Thm2} can be written also in the form \[ - \frac{\int_0^1 \cX_t\,\dd (\cX_t - mt)}{\int_0^1 \cX_t^2\,\dd t} = - \frac{\int_0^1 \cX_t\sqrt{\cY_t}\,\dd \cB_t} {\int_0^1 \cX_t^2\,\dd t}. \] (iii) By Proposition \ref{Pro_affine_2}, the affine process \ $(\cY_t,\cX_t)_{t\in\RR_+}$ \ given in Theorem \ref{Thm2} has infinitesimal generator \begin{align*} (\cA_{(\cY,\cX)} f)(x) = \frac{1}{2}x_1f_{1,1}''(x) + \frac{1}{2}x_1f_{2,2}''(x) + af_1'(x) + mf_2'(x) \end{align*} where \ $x=(x_1,x_2)\in\RR_+\times\RR$ \ and \ $f\in\cC^2_c(\RR_+\times\RR)$. \noindent(iv) Under the Condition (C), by Theorem \ref{Thm2} and Slutsky's lemma, we get \ $\widetilde\theta_n^{\mathrm{LSE}}$ \ converges stochastically to the parameter \ $\theta=0$ \ as \ $n\to\infty$.\proofend \end{Rem} \noindent{\bf Proof of Theorem \ref{Thm2}.} By \eqref{help22}, we have \[ X_t = X_0 + m t + \int_0^t \sqrt{Y_s} \, \dd B_s , \qquad t \in \RR_+ . \] Hence for all \ $t \in \RR_{++}$, \ the conditional distribution of \ $X_t$ \ given \ $X_0$ \ and \ $(Y_s)_{s\in[0,t]}$ \ is a normal distribution with mean \ $X_0 + mt$ \ and with variance \ $\int_0^t Y_s \, \dd s$. \ Here the variance \ $\int_0^t Y_s \, \dd s$ \ is positive almost surely for all \ $t \in \RR_{++}$. \ Indeed, let \ $A_t := \{ \omega \in \Omega : \text{$s \mapsto Y_s(\omega)$ \ is continuous on \ $[0, t]$} \}$. \ Then \ $\PP(A_t) = 1$, \ and, since \ $\PP(Y_0\geq 0)=1$, \ for all \ $\omega \in A_t$, \ $\int_0^t Y_s(\omega) \, \dd s = 0$ \ if and only if \ $Y_s(\omega) = 0$ \ for all \ $s \in [0, t]$. \ By \eqref{2dim_affine}, we have \[ Y_s = Y_0 + a s + \int_0^s \sqrt{Y_u} \, \dd W_u , \qquad s \in \RR_+ . \] The stochastic integral on the right hand side can be approximated as \[ \sup_{s\in[0,t]} \left| \sum_{i=1}^\ns \sqrt{Y_{(i-1)/n}} (W_{i/n} - W_{(i-1)/n}) - \int_0^s Y_u \, \dd W_u \right| \stoch 0 \qquad \text{as \ $n \to \infty$} \] for all \ $t\in\RR_+$, \ by Jacod and Shiryaev \cite[Theorem I.4.44]{JSh}. Hence there exists a sequence \ $(n_k)_{k\in\NN}$ \ of positive integers such that \[ \sup_{s\in[0,t]} \left| \sum_{i=1}^{\lfloor n_k t \rfloor} \sqrt{Y_{(i-1)/n_k}} (W_{i/n_k} - W_{(i-1)/n_k}) - \int_0^s Y_u \, \dd W_u \right| \as 0 \qquad \text{as \ $k \to \infty$} \] for all \ $t\in\RR_+$. \ Consequently, with the notation \[ \widetilde A_t := \left\{ \omega \in \Omega : \int_0^t Y_s(\omega) \, \dd s = 0 \right\}, \] we have \begin{align*} \widetilde A_t\cap A_t &\subset \left\{\widetilde A_t\bigcap\left\{\text{$\int_0^s Y_u \, \dd W_u = 0$ \ for all \ $s \in [0, t]$}\right\}\right\} \\ &\subset \left\{\widetilde A_t\cap \left\{\text{$Y_s = Y_0 + a s$ \ for all \ $s \in [0, t]$}\right\}\right\}\\ &=\left\{\widetilde A_t\cap \left\{\text{$Y_0s + a s^2/2=0$ \ for all \ $s \in [0, t]$}\right\}\right\}\\ &= \left\{\widetilde A_t\cap \left\{\text{$Y_0 = -a s/2$ \ for all \ $s \in [0, t]$}\right\}\right\} = \emptyset, \end{align*} implying \ $\PP\bigl(\int_0^t Y_s \, \dd s = 0\bigr) = 0$, \ and hence \ $\PP\bigl(\int_0^t Y_s \, \dd s > 0\bigr) = 1$. \ It yields that \begin{align}\label{help25} \PP(X_t = 0) = \EE\big(\PP(X_t = 0 \mid X_0, (Y_s)_{s\in[0,t]})\big) = 0, \qquad t \in \RR_{++} , \end{align} and hence \ $\PP(\sum_{i=1}^n X_{i-1}^2 > 0) = 1$ \ for all \ $n \geq 2$. Now we turn to prove \eqref{help23}. By It\^{o}'s formula, we have \ $\dd(\cX_t^2) = 2 \cX_t \dd \cX_t + \cY_t \, \dd t$, \ $t \in \RR_+$, \ and hence, using also \ $\cX_0 = 0$, \ we have \begin{align}\label{help49} \int_0^1 \cX_s \, \dd \cX_s = \frac{1}{2} \left( \cX_1^2 - \int_0^1 \cY_s \, \dd s \right) . \end{align} For the process \ $(X_t)_{t\in\RR_+}$, \ a discrete version \begin{align}\label{help53} \sum_{i=1}^n (X_i - X_{i-1}) X_{i-1} = \frac{1}{2} \left( X_n^2 - X_0^2 - \sum_{i=1}^n (X_i - X_{i-1})^2 \right) \end{align} of the identity \eqref{help49} can be easily checked. The aim of the following discussion is to prove \begin{align}\label{help10} \begin{split} &\biggl(\frac{1}{n^2} \sum_{i=1}^n X_{i-1} , \frac{1}{n^3} \sum_{i=1}^n X_{i-1}^2 , \frac{1}{n} X_n , \frac{1}{n} X_0 , \frac{1}{n^2} \sum_{i=1}^n (X_i - X_{i-1})^2 \biggr) \\ &\qquad\distr \left( \int_0^1 \cX_t \, \dd t , \int_0^1 \cX_t^2 \, \dd t , \cX_1 , 0, \int_0^1 \cY_t \, \dd t \right) \qquad \text{as \ $n\to\infty$.} \end{split} \end{align} Let us consider the unique strong solution of the SDE \begin{align}\label{help16_b} \begin{split} \begin{cases} \dd \tY_t = a \, \dd t + \sqrt{\tY_t} \, \dd W_t , \\[2mm] \dd \tX_t = m \, \dd t + \sqrt{\tY_t} \, \dd B_t , \end{cases}\qquad t \in \RR_+ , \end{split} \end{align} with initial value \ $(\tY_0, \tX_0)=(0, 0)$, \ where \ $(W_t)_{t\in\RR_+}$ \ and \ $(B_t)_{t\in\RR_+}$ \ are the independent standard \ $(\cF_t)_{t\in\RR_+}$-Wiener processes appearing in the SDE \eqref{2dim_affine}. First note that \ $(\tY_t, \tX_t)_{t\in\RR_+} \distre (\cY_t, \cX_t)_{t\in\RR_+}$, \ and, by Proposition \ref{Pro_affine_2}, it is an affine process having admissible parameters \[ \left( \begin{bmatrix} 0 & 0 \\ 0 & 0 \end{bmatrix} , \, \frac{1}{2} \begin{bmatrix} 1 & 0 \\ 0 & 1 \end{bmatrix} , \, \begin{bmatrix} a \\ m \end{bmatrix} , \, \begin{bmatrix} 0 & 0 \\ 0 & 0 \end{bmatrix} , \, 0 , 0 \right) , \] and condition \eqref{cond1} is trivially fulfilled. Hence, by part (ii) of Remark \ref{Rem1}, we have \begin{align}\label{help14} \left( n^{-1} \tY_{nt}, n^{-1} \tX_{nt} \right)_{t\in\RR_+} \distre (\cY_t, \cX_t)_{t\in\RR_+} \qquad \text{for all \ $n \in \NN$.} \end{align} Consequently, for all \ $n \in \NN$, \ we have \begin{align*} \biggl(\frac{1}{n^2} \sum_{i=1}^n \tX_{i-1} , \frac{1}{n^3} \sum_{i=1}^n \tX_{i-1}^2 , \frac{1}{n} \tX_n , \frac{1}{n} \tX_0 , \frac{1}{n^2} \sum_{i=1}^n (\tX_i - \tX_{i-1})^2 \biggr) \distre (A_n, B_n, C_n, D_n, E_n) , \end{align*} where \begin{align}\nonumber A_n &:= \frac{1}{n} \sum_{i=1}^n \cX_{(i-1)/n} \as \int_0^1 \cX_t \, \dd t , \qquad \text{as \ $n \to \infty$,} \\ \nonumber B_n &:= \frac{1}{n} \sum_{i=1}^n \cX_{(i-1)/n}^2 \as \int_0^1 \cX_t^2 \, \dd t , \qquad \text{as \ $n \to \infty$,} \\ \nonumber C_n &:= \cX_1 , \\ \nonumber D_n &:= \cX_0 , \\\label{help20} E_n &:= \sum_{i=1}^n (\cX_{i/n} - \cX_{(i-1)/n})^2 \stoch \int_0^1 \cY_t \, \dd t , \qquad \text{as \ $n \to \infty$.} \end{align} Here the first two convergences are consequences of the definition of the Riemann integral using also that \ $(\cX_t)_{t\in\RR_+}$ \ has continuous sample paths almost surely. The convergence \eqref{help20} can be checked as follows. With the notations of Jacod and Shiryaev \cite{JSh}, \ $\bigl(\tau_n:=\bigl(\frac{i}{n}\land 1\bigr)_{i\in\NN}\bigr)_{n\in\NN}$ \ is a Riemann sequence of (adapted) subdivisions and hence, by Jacod and Shiryaev \cite[Theorem I.4.47]{JSh}, the sequence of processes \[ \left(\sum_{i=1}^n \left(\cX\left(\frac{i}{n}\land 1\land t\right) -\cX\left(\frac{i-1}{n}\land 1\land t\right)\right)^2 \right)_{t\in\RR_+}, \qquad n\in\NN, \] converges to the quadratic variation process of \ $\cX$ \ in probability, uniformly on every compact interval. Especially, with \ $t=1$, \ using also the SDE \eqref{help16}, we have \eqref{help20}. Hence, in order to prove \eqref{help10}, it suffices to show convergences \begin{align} &\frac{1}{n^2} \sum_{i=0}^{n-1} X_i - \frac{1}{n^2} \sum_{i=0}^{n-1} \tX_i \stoch 0 , \label{conv1} \\ &\frac{1}{n^3} \sum_{i=0}^{n-1} X_i^2 - \frac{1}{n^3} \sum_{i=0}^{n-1} \tX_i^2 \stoch 0 , \label{conv2} \\ &\frac{1}{n} X_n - \frac{1}{n} \tX_n \stoch 0 , \label{conv3}\\ &\frac{1}{n} X_0 - \frac{1}{n} \tX_0 \stoch 0 , \label{conv4} \\ &\frac{1}{n^2} \sum_{i=1}^n (X_i - X_{i-1})^2 - \frac{1}{n^2} \sum_{i=1}^n (\tX_i - \tX_{i-1})^2 \stoch 0 , \label{conv5} \end{align} as \ $n \to \infty$. \ Indeed, one can refer to Slutsky's lemma using also that for any random vectors \ $U_n$, \ $V_n$, \ $n\in\NN$, \ $U$, \ $V$ \ such that \ $U_n \stoch U$ \ and \ $V_n \stoch V$ \ as \ $n \to \infty$, we have \ $(U_n,V_n) \stoch (U,V)$ \ as \ $n \to \infty$, \ see, e.g., van der Vaart \cite[Theorem 2.7, part (vi)]{Vaa}. The convergence \eqref{conv4} is trivial. Next we show \begin{equation}\label{EY-tY} \EE(|Y_t - \tY_t|) \leq \EE(Y_0) , \qquad t \in \RR_+ . \end{equation} By \eqref{2dim_affine} and \eqref{help16_b}, we have \[ Y_t - \tY_t = Y_0 + \int_0^t (\sqrt{Y_s} - \sqrt{\tY_s}) \, \dd W_s , \qquad t \in \RR_+ . \] For each \ $n \in \NN$, \ there exists an even and twice continuously differentiable function \ $\psi_n : \RR \to \RR_+$ \ with \ $\vert\psi_n(x)\vert\leq \vert x\vert$, \ $|\psi_n'(x)| \leq 1$, \ $\psi_n(x) \uparrow |x|$ \ as \ $n \to \infty$ \ for all \ $x \in \RR$, \ and \begin{align*} \psi_n''(x-y)(\sqrt{x} - \sqrt{y})^2 \leq \frac{2(\sqrt{x} - \sqrt{y})^2}{n\vert x-y\vert} \leq \frac{2}{n} \end{align*} for all \ $n\in\NN$ \ and \ $x,y\in\RR_+$, \ see, e.g., in Karatzas and Shreve \cite[Proof of Proposition 5.2.13]{KarShr}. By It\^o's formula, \begin{align}\label{help51} \begin{split} \psi_n(Y_t - \tY_t) &= \psi_n(Y_0) + \frac{1}{2} \int_0^t \psi_n''(Y_s - \tY_s) \left(\sqrt{Y_s} - \sqrt{\tY_s}\right)^2 \, \dd s\\ &\phantom{=\;} + \int_0^t \psi_n'(Y_s - \tY_s) \left(\sqrt{Y_s} - \sqrt{\tY_s}\right) \, \dd W_s \end{split} \end{align} for all \ $t \in \RR_+$ \ and \ $n \in \NN$. \ The last term is an \ $(\cF_t)_{t\in\RR_+}$-martingale, since \begin{align*} \EE\left( \int_0^t |\psi_n'(Y_s - \tY_s)| \left(\sqrt{Y_s} - \sqrt{\tY_s}\right)^2 \, \dd s \right) &\leq \EE\left( \int_0^t |Y_s - \tY_s| \, \dd s \right)\\ &\leq \int_0^t (\EE(Y_s) + \EE(\tY_s)) \, \dd s < \infty, \end{align*} where the last inequality follows by Lemma \ref{Pro_moments}. Thus the expectation of the last term on the right hand side of \eqref{help51} is zero, whereas the expectation of the second integral is bounded by \ $2t/n$. \ We conclude \[ \EE(\psi_n(Y_t - \tY_t)) \leq \EE(\psi_n(Y_0)) + \frac{t}{n} , \qquad t \in \RR_+ , \quad n \in \NN . \] By monotone convergence theorem, we have \begin{align*} &\EE(\vert Y_t - \tY_t\vert) = \EE(\lim_{n\to\infty} \psi_n(Y_t - \tY_t) ) = \lim_{n\to\infty} \EE(\psi_n(Y_t - \tY_t))\\ &\quad \leq \liminf_{n\to\infty} \left(\EE(\psi_n(Y_0)) + \frac{t}{n}\right) = \lim_{n\to\infty} \EE(\psi_n(Y_0)) = \EE(\lim_{n\to\infty} \psi_n(Y_0)) =\EE(\vert Y_0\vert), \end{align*} which yields \eqref{EY-tY}. Next, we derive \begin{equation}\label{EX-tX} \EE(|X_t - \tX_t|) \leq \EE(|X_0|) + \sqrt{t \EE(Y_0)} , \qquad t \in \RR_+ . \end{equation} Again by \eqref{2dim_affine} and \eqref{help16_b}, we have \begin{align}\label{help52} X_t - \tX_t = X_0 + \int_0^t \left(\sqrt{Y_s} - \sqrt{\tY_s}\right) \, \dd B_s , \qquad t \in \RR_+ , \end{align} hence \begin{align*} \EE(|X_t - \tX_t|) &\leq \EE(|X_0|) + \sqrt{\EE\left(\left(\int_0^t (\sqrt{Y_s} - \sqrt{\tY_s}) \, \dd B_s\right)^2\right)} \\ &= \EE(|X_0|) + \sqrt{\EE\left(\int_0^t \left(\sqrt{Y_s} - \sqrt{\tY_s}\right)^2 \, \dd s\right)}\\ & \leq \EE(|X_0|) + \sqrt{\EE\left(\int_0^t |Y_s - \tY_s| \, \dd s\right)} \\ &= \EE(|X_0|) + \sqrt{\int_0^t \EE(|Y_s - \tY_s|) \, \dd s} , \end{align*} which yields \eqref{EX-tX} by \eqref{EY-tY}. By \eqref{EX-tX}, we have \[ \EE\left(\left|\frac{1}{n} X_n - \frac{1}{n} \tX_n\right|\right) \leq \frac{1}{n} \left( \EE(|X_0|) + \sqrt{n \EE(Y_0)} \right) \to 0 \qquad \text{as \ $n \to \infty$,} \] hence we obtain \eqref{conv3}. In a similar way, \[ \EE\left(\left| \frac{1}{n^2} \sum_{i=0}^{n-1} X_i - \frac{1}{n^2} \sum_{i=0}^{n-1} \tX_i \right|\right) \leq \frac{1}{n^2} \sum_{i=0}^{n-1} \left( \EE(|X_0|) + \sqrt{i \EE(Y_0)} \right) \to 0 \] as \ $n \to \infty$, \ hence we obtain \eqref{conv1}. We also have \begin{equation}\label{EX-tX2} \EE\left((X_t - \tX_t)^2\right) \leq 2 \EE(X_0^2) + 2 t \EE(Y_0) , \qquad t \in \RR_+ . \end{equation} Indeed, by \eqref{help52}, using Minkowski inequality, we have \begin{align*} \sqrt{\EE\left((X_t - \tX_t)^2\right)} &\leq \sqrt{\EE(X_0^2)} + \sqrt{\EE\left(\left(\int_0^t (\sqrt{Y_s} - \sqrt{\tY_s}) \, \dd B_s\right)^2\right)} \\ &= \sqrt{\EE(X_0^2)} + \sqrt{\EE\left(\int_0^t \left(\sqrt{Y_s} - \sqrt{\tY_s}\right)^2 \, \dd s\right)}\\ & \leq \sqrt{\EE(X_0^2)} + \sqrt{\EE\left(\int_0^t |Y_s - \tY_s| \, \dd s\right)} \\ &= \sqrt{\EE(X_0^2)} + \sqrt{\int_0^t \EE(|Y_s - \tY_s|) \, \dd s} \leq \sqrt{\EE(X_0^2)} + \sqrt{t \EE(Y_0)} \end{align*} by \eqref{EY-tY}, which yields \eqref{EX-tX2}. In a similar way, \begin{equation}\label{EX2} \EE(X_t^2) \leq 3 \EE(X_0^2) + 3 m^2t^2 + 3 t \EE(Y_0) + 3 a t^2 /2 , \qquad t \in \RR_+ . \end{equation} since, by \eqref{2dim_affine} and \eqref{Pro_moments}, \begin{align*} \sqrt{\EE(X_t^2)} &\leq \sqrt{\EE(X_0^2)} + |m| t + \sqrt{\EE\left(\left(\int_0^t \sqrt{Y_s} \, \dd B_s\right)^2\right)}\\ & = \sqrt{\EE(X_0^2)} + |m| t + \sqrt{\EE\left(\int_0^t Y_s \, \dd s\right)} \\ &= \sqrt{\EE(X_0^2)} + |m| t + \sqrt{\int_0^t (\EE(Y_0) + a s) \, \dd s}\\ &= \sqrt{\EE(X_0^2)} + |m| t + \sqrt{t \EE(Y_0) + a t^2 /2} \end{align*} for \ $t\in\RR_+$, \ which yields \eqref{EX2}. Clearly, \eqref{EX2} implies also \ $\EE(\tX_t^2) \leq 3m^2t^2 + 3 a t^2 /2$ \ for all \ $t \in \RR_+$, \ and hence, together with \eqref{EX-tX2} and \eqref{EX2}, we conclude \begin{align*} &\EE\left(\left| \frac{1}{n^3} \sum_{i=0}^{n-1} X_i^2 - \frac{1}{n^3} \sum_{i=0}^{n-1} \tX_i^2 \right|\right) \leq \frac{1}{n^3} \sum_{i=0}^{n-1} \EE\left(|(X_i - \tX_i) (X_i + \tX_i)|\right) \\ &\quad\leq \frac{1}{n^3} \sum_{i=0}^{n-1} \sqrt{ \EE((X_i - \tX_i)^2) \EE((X_i + \tX_i)^2)} \\ &\quad\leq \frac{1}{n^3} \sum_{i=0}^{n-1} \sqrt{ 2 \EE((X_i - \tX_i)^2) (\EE(X_i^2) + \EE(\tX_i^2))} \\ &\quad\leq \frac{1}{n^3} \sum_{i=0}^{n-1} \sqrt{ 12 (\EE(X_0^2) + i \EE(Y_0)) \left(\EE(X_0^2) + (\EE(Y_0)) i + (2m^2 + a)i^2 \right)} \to 0 \end{align*} as \ $n \to \infty$, \ hence we obtain \eqref{conv2}. Next, we show \eqref{conv5}. Again by \eqref{2dim_affine} and \eqref{help16_b}, we have \[ X_i - X_{i-1} = m + \int_{i-1}^i \sqrt{Y_s} \, \dd B_s , \qquad \tX_i - \tX_{i-1} = m + \int_{i-1}^i \sqrt{\tY_s} \, \dd B_s , \qquad i \in \NN , \] and hence \begin{align*} &\frac{1}{n^2} \sum_{i=1}^n (X_i - X_{i-1})^2 - \frac{1}{n^2} \sum_{i=1}^n (\tX_i - \tX_{i-1})^2 \\ &\qquad= \frac{2m}{n^2} \int_0^n \left(\sqrt{Y_s} - \sqrt{\tY_s}\right) \, \dd B_s\\ &\qquad\phantom{=\;} + \frac{1}{n^2} \sum_{i=1}^n \left[\left(\int_{i-1}^i \sqrt{Y_s} \, \dd B_s\right)^2 - \left(\int_{i-1}^i \sqrt{\tY_s} \, \dd B_s\right)^2\right]\\ &\qquad=: 2m R_n + S_n . \end{align*} Here, by \eqref{EY-tY}, \begin{align*} \EE(R_n^2) &= \frac{1}{n^4} \EE\left(\left(\int_0^n \left(\sqrt{Y_s} - \sqrt{\tY_s}\right) \, \dd B_s\right)^2\right)\\ & = \frac{1}{n^4} \EE\left(\int_0^n \left(\sqrt{Y_s} - \sqrt{\tY_s}\right)^2 \, \dd s\right) \leq \frac{1}{n^4} \EE\left(\int_0^n |Y_s - \tY_s| \, \dd s\right)\\ & = \frac{1}{n^4} \int_0^n \EE(|Y_s - \tY_s|) \, \dd s \leq \frac{\EE(Y_0)}{n^3} \to 0 \qquad \text{as \ $n \to \infty$,} \end{align*} hence \ $R_n \stoch 0$ \ as \ $n \to \infty$. \ Further, by \eqref{EY-tY}, \begin{align*} &\EE(|S_n|) = \EE\left(\left| \frac{1}{n^2} \sum_{i=1}^n \int_{i-1}^i \left(\sqrt{Y_s} - \sqrt{\tY_s}\right) \, \dd B_s \int_{i-1}^i \left(\sqrt{Y_s} + \sqrt{\tY_s}\right) \, \dd B_s \right|\right) \\ &\leq \frac{1}{n^2} \sum_{i=1}^n \EE\left(\left| \int_{i-1}^i \left(\sqrt{Y_s} - \sqrt{\tY_s}\right) \, \dd B_s \int_{i-1}^i \left(\sqrt{Y_s} + \sqrt{\tY_s}\right) \, \dd B_s \right|\right) \\ &\leq \frac{1}{n^2} \sum_{i=1}^n \sqrt{\EE\left(\left( \int_{i-1}^i (\sqrt{Y_s} - \sqrt{\tY_s}) \, \dd B_s\right)^2\right) \EE\left(\left( \int_{i-1}^i (\sqrt{Y_s} + \sqrt{\tY_s}) \, \dd B_s\right)^2\right)} \\ &= \frac{1}{n^2} \sum_{i=1}^n \sqrt{\int_{i-1}^i \EE\left((\sqrt{Y_s} - \sqrt{\tY_s})^2\right) \, \dd s \int_{i-1}^i \EE\left((\sqrt{Y_s} + \sqrt{\tY_s})^2\right) \, \dd s} \\ &\leq \frac{1}{n^2} \sum_{i=1}^n \sqrt{\int_{i-1}^i \EE(|Y_s - \tY_s|) \, \dd s \int_{i-1}^i 2 (\EE(Y_s) + \EE(\tY_s)) \, \dd s}\\ &\leq \frac{1}{n^2} \sum_{i=1}^n \sqrt{\EE(Y_0) \int_{i-1}^i 2(\EE(Y_0) + 2 a s) \, \dd s} \\ &= \frac{1}{n^2} \sum_{i=1}^n \sqrt{2 \EE(Y_0) (\EE(Y_0) + (2 i - 1) a)} \to 0 \qquad \text{as \ $n \to \infty$,} \end{align*} thus \ $S_n \stoch 0$ \ as \ $n \to \infty$, \ and we obtain \eqref{conv5}, and hence \eqref{help10}. Finally, by \eqref{help10} and the continuous mapping theorem, and using that \begin{align*} n\widetilde\theta_n^{\mathrm{LSE}} = \frac{\frac{m}{n^2} \sum_{i=1}^n X_{i-1} - \frac{1}{2n^2} X_n^2 + \frac{1}{2n^2} X_0^2 + \frac{1}{2n^2} \sum_{i=1}^n (X_i - X_{i-1})^2} {\frac{1}{n^3} \sum_{i=1}^n X_{i-1}^2} , \end{align*} we have the assertion. Indeed, the function \ $g : \RR^5 \to \RR$, \ defined by \begin{align*} g(x, y, z, u, v) :=\begin{cases} \frac{m x - (z^2 - u^2 - v)/2}{y} & \text{if \ $y \ne 0$,} \\ 0 & \text{if \ $y = 0$,} \end{cases} \end{align*} is continuous on the set \ $\{(x, y, z, u, v) \in \RR^5 : y \ne 0\}$, \ and the limit distribution in \eqref{help10} is concentrated on this set since \ $\PP\bigl(\int_0^1 \cX_t^2 \, \dd t > 0\bigr) = 1$. \ Indeed, if \ $\PP(\int_0^1\cX_t^2\,\dd t =0 )>0$ \ held, then, by the almost sure continuity of the sample paths of \ $\cX$, \ we would have \ $\PP(\cX_t = 0,\; \forall\, t\in[0,1])>0$. \ Hence on the event \ $\{\omega\in\Omega : \cX_t(\omega) = 0, \; \forall\, t\in[0,1]\}$, \ the quadratic variation of \ $\cX$ \ would be identically zero. Since \ $\dd \cX_t = m\,\dd t + \sqrt{\cY_t}\,\dd \cB_t$, $t\in\RR_+$, \ the quadratic variation of \ $\cX$ \ is the process \ $\bigl(\int_0^t\cY_u\,\dd u\bigr)_{t\in\RR_+}$, \ and then we would have \[ \int_0^t \cY_u\,\dd u = 0 \quad \text{for all \ $t\in[0,1]$} \] on the event \ $\{\omega\in\Omega : \cX_t(\omega) = 0, \; \forall\, t\in[0,1]\}$. \ This yields us to a contradiction similarly as at the beginning of the proof due to that \ $a\in\RR_{++}$ \ and \ $\dd\cY_t = a\,\dd t + \sqrt{\cY_t}\,\dd\cW_t$, $t\in\RR_+$. \ Hence the continuous mapping theorem (see, e.g., Theorem 2.3 in van der Vaart \cite{Vaa}) yields \begin{multline*} g\left( \frac{1}{n^2} \sum_{i=1}^n X_{i-1} , \frac{1}{n^3} \sum_{i=1}^n X_{i-1}^2 , \frac{1}{n} X_n , \frac{1}{n} X_0 , \frac{1}{n^2} \sum_{i=1}^n (X_i-X_{i-1})^2 \right) \\ \distr g\left( \int_0^1 \cX_t \, \dd t , \int_0^1 \cX_t^2 \, \dd t , \cX_1 , 0 , \int_0^1 \cY_t \, \dd t \right) \qquad \text{as \ $n \to \infty$.} \end{multline*} \ Since \begin{align*} &\PP\left( n\widetilde\theta_n^{\mathrm{LSE}} = g\left( \frac{1}{n^2} \sum_{i=1}^n X_{i-1} , \frac{1}{n^3} \sum_{i=1}^n X_{i-1}^2 , \frac{1}{n} X_n , \frac{1}{n} X_0 , \frac{1}{n^2} \sum_{i=1}^n (X_i - X_{i-1})^2 \right) \right)\\ &\qquad\geq \PP\left(\sum_{i=1}^n X_{i-1}^2 > 0 \right) = 1 \end{align*} for all \ $n \geq 2$, \ the assertion follows using \eqref{help49} and that if \ $\xi_n,$ \ $\eta_n$, \ $n \in \NN$, \ and \ $\xi$ \ are random variables such that \ $\xi_n \distr \xi$ \ as \ $n \to \infty$ \ and \ $\lim_{n\to\infty} \PP(\xi_n = \eta_n) = 1$, \ then \ $\eta_n \distr \xi$ \ as \ $n \to \infty$, \ see, e.g., Barczy et al.\ \cite[Lemma 3.1]{BarIspPap1}. \proofend \begin{Rem}\label{Rem3} If the affine diffusion process given by the SDE \eqref{2dim_affine} is critical but \ $(b,\theta)\ne(0,0)$ \ (i.e., \ $b=0$, $\theta>0$ \ or \ $b>0$, $\theta=0$), \ then the asymptotic behaviour of the LSE \ $\widetilde\theta_n^{\mathrm{LSE}}$ \ cannot be studied using Theorem \ref{Thm1} since its condition \ $\lim_{\theta\to\infty} \theta \beta^{(\theta)} = \beta$ \ is not satisfied. \proofend \end{Rem} \subsection{Least squares estimator of \ $(\theta,m)$} \label{section_LSE_M_theta} The LSE of \ $(\theta,m)$ \ based on the observations \ $X_i$, \ $i=0,1,\ldots,n$, \ can be obtained by solving the extremum problem \[ (\widehat\theta_n^{\mathrm{LSE}}, \widehat m_n^{\mathrm{LSE}}) := \argmin_{(\theta,m)\in\RR^2} \sum_{i=1}^n (X_i - X_{i-1} - (m-\theta X_{i-1}))^2. \] We need to solve the following system of equations with respect to \ $(\theta,m)$: \begin{align*} &2\sum_{i=1}^n (X_i - X_{i-1} - (m-\theta X_{i-1})) X_{i-1} = 0,\\ &2\sum_{i=1}^n (X_i - X_{i-1} - (m-\theta X_{i-1})) = 0, \end{align*} which can be written also in the form \[ \begin{bmatrix} \sum_{i=1}^n X_{i-1}^2 & - \sum_{i=1}^n X_{i-1} \\ - \sum_{i=1}^n X_{i-1} & n \end{bmatrix} \begin{bmatrix} \theta \\ m \end{bmatrix} = \begin{bmatrix} - \sum_{i=1}^n (X_i - X_{i-1}) X_{i-1} \\ \sum_{i=1}^n (X_i - X_{i-1}) \end{bmatrix}. \] Then one can check that \begin{align}\label{help12} \widehat\theta_n^{\mathrm{LSE}} = - \frac{n\sum_{i=1}^n (X_i - X_{i-1})X_{i-1} - \sum_{i=1}^n X_{i-1} \sum_{i=1}^n (X_i-X_{i-1})} {n\sum_{i=1}^n X_{i-1}^2 - \left(\sum_{i=1}^n X_{i-1}\right)^2}, \end{align} and \begin{align}\label{help13} \widehat m_n^{\mathrm{LSE}} = \frac{\sum_{i=1}^n X_{i-1}^2 \sum_{i=1}^n (X_i-X_{i-1}) - \sum_{i=1}^n X_{i-1} \sum_{i=1}^n (X_i-X_{i-1})X_{i-1}} {n\sum_{i=1}^n X_{i-1}^2 - \left(\sum_{i=1}^n X_{i-1}\right)^2} \end{align} provided that \ $n\sum_{i=1}^n X_{i-1}^2 - \left(\sum_{i=1}^n X_{i-1}\right)^2>0$. \ Since the matrix \[ \begin{bmatrix} 2\sum_{i=1}^n X_{i-1}^2 & -2\sum_{i=1}^n X_{i-1} \\ -2\sum_{i=1}^n X_{i-1} & 2n \end{bmatrix} \] which consists of the second order partial derivatives of the function $\RR^2\ni(\theta,m)\mapsto\sum_{i=1}^n (X_i - X_{i-1} - (m-\theta X_{i-1}))^2$ \ is positive definite provided that \ $n\sum_{i=1}^n X_{i-1}^2 - \left(\sum_{i=1}^n X_{i-1}\right)^2>0$, \ we have \ $(\widehat\theta_n^{\mathrm{LSE}},\widehat m_n^{\mathrm{LSE}})$ \ is indeed the solution of the extremum problem provided that \ $n\sum_{i=1}^n X_{i-1}^2 - \left(\sum_{i=1}^n X_{i-1}\right)^2>0$. \begin{Thm}\label{Thm3} Let us assume that Condition (C) holds. Then \begin{align}\label{help26} \PP\left(n\sum_{i=1}^n X_{i-1}^2 - \left(\sum_{i=1}^n X_{i-1}\right)^2>0\right)=1 \qquad \text{for all \ $n\geq 2$,} \end{align} and there exists a unique LSE \ $(\widehat\theta_n^{\mathrm{LSE}},\widehat m_n^{\mathrm{LSE}})$ \ which has the form given in \eqref{help12} and \eqref{help13}. Further, \begin{align*} n\widehat\theta_n^{\mathrm{LSE}} \distr - \frac{\int_0^1 \cX_t\,\dd \cX_t - \cX_1\int_0^1 \cX_t\,\dd t} {\int_0^1 \cX_t^2\,\dd t - \left(\int_0^1 \cX_t\,\dd t\right)^2} \qquad \text{as \ $n\to\infty$,} \end{align*} and \begin{align*} \widehat m_n^{\mathrm{LSE}} \distr \frac{\cX_1\int_0^1 \cX_t^2\,\dd t - \int_0^1 \cX_t\,\dd t \int_0^1 \cX_t\,\dd \cX_t} {\int_0^1 \cX_t^2\,\dd t - \left(\int_0^1 \cX_t\,\dd t\right)^2} \qquad \text{as \ $n\to\infty$,} \end{align*} where \ $(\cX_t)_{t\in\RR_+}$ \ is the second coordinate of a two-dimensional affine process\ $(\cY_t,\cX_t)_{t\in\RR_+}$ \ given by the unique strong solution of the SDE \begin{align*} \begin{cases} \dd\cY_t = a\,\dd t + \sqrt{\cY_t}\,\dd \cW_t,\\[2mm] \dd\cX_t = m\,\dd t + \sqrt{\cY_t}\,\dd \cB_t, \end{cases} \qquad t\in\RR_+, \end{align*} with initial value \ $(\cY_0,\cX_0)=(0,0)$, \ where \ $(\cW_t)_{t\in\RR_+}$ \ and \ $(\cB_t)_{t\in\RR_+}$ \ are independent standard Wiener processes. \end{Thm} \begin{Rem}\label{Rem6} (i) The limit distributions in Theorem \ref{Thm3} have the same forms as those of the limit distributions in Theorem 4.1 in Hu and Long \cite{HuLon3}. \noindent (ii) By Proposition \ref{Pro_affine_2}, the affine process \ $(\cY_t,\cX_t)_{t\in\RR_+}$ \ given in Theorem \ref{Thm3} has infinitesimal generator \begin{align*} (\cA_{(\cY,\cX)} f)(x) = \frac{1}{2}x_1f_{1,1}''(x) + \frac{1}{2}x_1f_{2,2}''(x) + af_1'(x) + mf_2'(x), \end{align*} where \ $x=(x_1,x_2)\in\RR_+\times\RR$ \ and \ $f\in\cC^2_c(\RR_+\times\RR)$. \noindent(iii) Under the Condition (C), by Theorem \ref{Thm3} and Slutsky's lemma, we get \ $\widehat\theta_n^{\mathrm{LSE}}$ \ converges stochastically to the parameter \ $\theta=0$ \ as \ $n\to\infty$, \ and one can show that \ $\widehat m_n^{\mathrm{LSE}}$ \ does not converge stochastically to the parameter \ $m$ \ as \ $n\to\infty$, \ see Appendix \ref{AppendixB}. \proofend \end{Rem} \noindent{\bf Proof of Theorem \ref{Thm3}.} By an easy calculation, \begin{align*} n\sum_{i=1}^n X_{i-1}^2 - \left(\sum_{i=1}^n X_{i-1}\right)^2 = n \sum_{i=1}^n \left( X_{i-1} - \frac{1}{n}\sum_{j=1}^n X_{j-1} \right)^2 \geq 0, \end{align*} and equality holds if and only if \[ X_{i-1} = \frac{1}{n} \sum_{j=1}^n X_{j-1},\quad i=1,\ldots,n \qquad \Longleftrightarrow \qquad X_0=X_1=\cdots=X_{n-1}. \] By \eqref{help22}, for all \ $n\geq 2$, \begin{align*} \PP(X_0=X_1=\cdots=X_{n-1}) &\leq \PP(X_0 = X_1) = \PP\left(\int_0^1\sqrt{Y_s}\,\dd B_s = m\right)\\ &= \EE\left(\PP\left( \int_0^1\sqrt{Y_s}\,\dd B_s = m \;\Big\vert\; (Y_s)_{s\in[0,1]}\right)\right)=0, \end{align*} where we used that the conditional distribution of \ $\int_0^1\sqrt{Y_s}\,\dd B_s$ \ given \ $(Y_s)_{s\in[0,1]}$ \ is a normal distribution with mean \ $0$ \ and with variance \ $\int_0^1 Y_s\,\dd s$. \ Here the variance \ $\int_0^1 Y_s\,\dd s$ \ is positive almost surely, see the proof of Theorem \ref{Thm2}. This yields \eqref{help26}. By \eqref{help12} and \eqref{help13}, we have \begin{align*} &n\widehat\theta_n^{\mathrm{LSE}} = - \frac{\frac{1}{n^2}\sum_{i=1}^n (X_i - X_{i-1})X_{i-1} - \frac{1}{n^2} \sum_{i=1}^n X_{i-1} \frac{1}{n}(X_n-X_0)} {\frac{1}{n^3}\sum_{i=1}^n X_{i-1}^2 - \left(\frac{1}{n^2}\sum_{i=1}^n X_{i-1}\right)^2},\\ &\widehat m_n^{\mathrm{LSE}} = \frac{\frac{1}{n^3}\sum_{i=1}^n X_{i-1}^2 \frac{1}{n}(X_n-X_0) - \frac{1}{n^2}\sum_{i=1}^n X_{i-1} \frac{1}{n^2}\sum_{i=1}^n (X_i - X_{i-1})} {\frac{1}{n^3}\sum_{i=1}^n X_{i-1}^2 - \left(\frac{1}{n^2}\sum_{i=1}^n X_{i-1}\right)^2}, \end{align*} and using \eqref{help53} and \eqref{help10}, as in the proof of Theorem \ref{Thm2}, the continuous mapping theorem yields the assertion. We only remark that \begin{align}\label{help37} \PP\left(\int_0^1 \cX_t^2\,\dd t - \left(\int_0^1 \cX_t\,\dd t\right)^2 > 0\right) = 1. \end{align} Indeed, \begin{align*} \int_0^1 \cX_t^2\,\dd t - \left(\int_0^1 \cX_t\,\dd t\right)^2 = \int_0^1 \left(\cX_t - \int_0^1 \cX_s\,\dd s\right)^2 \geq 0, \end{align*} and equality holds if and only if \[ \cX_t = \int_0^1 \cX_s\,\dd s \quad \text{a.e. \ $t\in[0,1]$.} \] Since \ $\cX$ \ has continuous sample paths almost surely, \begin{align}\label{help41} \PP\left( \int_0^1 \cX_t^2\,\dd t - \left(\int_0^1 \cX_t\,\dd t\right)^2 = 0 \right)>0 \end{align} holds if and only if \[ \PP\left( \cX_t = \int_0^1 \cX_s\,\dd s,\; \forall\,t\in[0,1] \right)>0. \] Hence, since \ $\cX_0=0$, \ we have \eqref{help41} holds if and only if \ $\PP(\cX_t = 0,\;\forall\,t\in[0,1] )>0$, \ which is a contradiction due to our assumption \ $a\in\RR_{++}$ \ (for more details, see the proof of Theorem \ref{Thm2}). \proofend \subsection{Conditional least squares estimator of \ $(\theta,m)$} For all \ $t\in\RR_+$, \ let \ $\cF^{(Y,X)}_t$ \ be the \ $\sigma$-algebra generated by \ $(Y_s,X_s)_{s\in[0,t]}$. \ The conditional least squares estimator (CLSE) of \ $(\theta,m)$ \ based on the observations \ $X_i$, $i=0,1,\ldots,n$, \ can be obtained by solving the extremum problem \[ (\widehat\theta_n^{\mathrm {CLSE}},\widehat m_n^{\mathrm{CLSE}}) := \argmin_{(\theta,m)\in\RR^2} \sum_{i=1}^n \Big(X_i - \EE(X_i\mid \cF_{i-1}^{(Y,X)})\Big)^2. \] By \eqref{help22}, for all \ $(y_0,x_0)\in\RR_+\times\RR$, \ we have \begin{align*} \EE\big( X_t \mid (Y_0,X_0) = (y_0,x_0)\big) = \ee^{-\theta t}x_0 + m\int_0^t\ee^{-\theta(t-u)}\,\dd u, \qquad t\geq 0, \end{align*} where we used that \ $\bigl( \int_0^t\ee^{\theta u}\sqrt{Y_u}\,\dd B_u \bigr)_{t\in\RR_+}$ \ is a martingale (which follows by the proof of Proposition \ref{Pro_moments}). Using that \ $(Y_t,X_t)_{t\in\RR_+}$ \ is a time-homogeneous Markov process, we have \begin{align*} \EE( X_t \mid \cF^{(Y,X)}_s) = \EE( X_t \mid (Y_s,X_s)) = \ee^{-\theta(t-s)}X_s + m\int_s^t\ee^{-\theta(t-u)}\,\dd u \end{align*} for \ $0\leq s\leq t$. \ Then \begin{align*} X_i - \EE( X_i \mid \cF^{(Y,X)}_{i-1}) & = X_i - \ee^{-\theta} X_{i-1} - m \int_{i-1}^i\ee^{-\theta(i-u)}\,\dd u\\ & = X_i - \ee^{-\theta} X_{i-1} - m \int_0^1\ee^{-\theta v}\,\dd v\\ & = X_i - \gamma X_{i-1} - \delta, \qquad i=1,\ldots,n, \end{align*} where \[ \gamma:= \ee^{-\theta} \qquad \text{and} \qquad \delta:=m\int_0^1\ee^{-\theta v}\,\dd v =\begin{cases} m\frac{1-\ee^{-\theta}}{\theta} & \text{if \ $\theta\ne0$,}\\ m & \text{if \ $\theta=0$.} \end{cases} \] Hence for all \ $n\in\NN$, \begin{align}\label{help29} \begin{split} & \widehat\gamma_n^{\mathrm{CLSE}} = \ee^{-\widehat\theta_n^{\mathrm{CLSE}}},\\ & \widehat\delta_n^{\mathrm{CLSE}} = \widehat m_n^{\mathrm{CLSE}} \int_0^1\ee^{-\widehat \theta_n^{\mathrm{CLSE}} v}\,\dd v, \end{split} \end{align} where \ $(\widehat\gamma_n^{\mathrm{CLSE}}, \widehat\delta_n^{\mathrm{CLSE}})$ \ is a CLSE of \ $(\gamma,\delta)$ \ based on the observations \ $X_i$, $i=0,1,\ldots,n$, \ which can be obtained by solving the extremum problem \begin{align}\label{help40} (\widehat\gamma_n^{\mathrm{CLSE}},\widehat\delta_n^{\mathrm{CLSE}}) := \argmin_{(\gamma,\delta)\in\RR^2} \sum_{i=1}^n (X_i - \gamma X_{i-1} - \delta)^2. \end{align} Indeed, the function \ $A:\RR^2\to\RR^2$, \[ \RR^2\ni (\theta',m') \mapsto A(\theta',m') := \begin{bmatrix} \gamma'\\ \delta' \\ \end{bmatrix} =: \begin{bmatrix} \ee^{-\theta'} \\ m'\int_0^1\ee^{-\theta'v}\,\dd v \\ \end{bmatrix} \in \RR_+\times\RR \] is bijective and measurable, and then there is a bijection between the set of CLSEs of the parameters \ $(\theta,m)$ \ and the set\vadjust{\vfill\eject} of CLSEs of the parameters \ $A(\theta,m)$. \ This follows easily, since for all \ $n\in\NN$, \ $(x_0,x_1,\ldots,x_n)\in\RR^{n+1}$ \ and \ $(\gamma', \delta')\in\RR_+\times\RR$, \begin{align*} \sum_{i=1}^n (x_i - \gamma' x_{i-1} - \delta')^2 = \sum_{i=1}^n \left(x_i - \begin{bmatrix} \gamma' \\ \delta' \\ \end{bmatrix}^\top \begin{bmatrix} x_{i-1} \\ 1 \\ \end{bmatrix} \right)^2 = \sum_{i=1}^n \left(x_i - \left(A(\theta',m')\right)^\top \begin{bmatrix} x_{i-1} \\ 1 \\ \end{bmatrix} \right)^2, \end{align*} hence \ $(\widehat\theta_n^{\mathrm{CLSE}},\widehat m_n^{\mathrm{CLSE}})$ \ is a CLSE of \ $(\theta,m)$ \ if and only if \ $A(\widehat\theta_n^{\mathrm{CLSE}},\widehat m_n^{\mathrm{CLSE}})$ \ is a CLSE of \ $A(\theta,m)$. For the extremum problem \eqref{help40}, we need to solve the following system of equations with respect to \ $(\gamma,\delta)$: \begin{align*} &2\sum_{i=1}^n ( X_i - \gamma X_{i-1} - \delta ) X_{i-1} = 0,\\ &2\sum_{i=1}^n ( X_i - \gamma X_{i-1} - \delta ) = 0, \end{align*} which can be written also in the form \[ \begin{bmatrix} \sum_{i=1}^n X_{i-1}^2 & \sum_{i=1}^n X_{i-1} \\ \sum_{i=1}^n X_{i-1} & n \\ \end{bmatrix} \begin{bmatrix} \gamma \\ \delta \\ \end{bmatrix} = \begin{bmatrix} \sum_{i=1}^n X_{i-1}X_i \\ \sum_{i=1}^n X_i \\ \end{bmatrix}. \] Then \begin{align}\label{help43} \widehat\gamma_n^{\mathrm{CLSE}} = \frac{n\sum_{i=1}^n X_{i-1}X_i - \sum_{i=1}^n X_{i-1} \sum_{i=1}^n X_i} {n\sum_{i=1}^n X_{i-1}^2 - \left(\sum_{i=1}^n X_{i-1}\right)^2}, \end{align} and \begin{align}\label{help44} \widehat \delta_n^{\mathrm{CLSE}} = \frac{\sum_{i=1}^n X_{i-1}^2 \sum_{i=1}^n X_i - \sum_{i=1}^n X_{i-1} \sum_{i=1}^n X_{i-1}X_i} {n\sum_{i=1}^n X_{i-1}^2 - \left(\sum_{i=1}^n X_{i-1}\right)^2}, \end{align} provided that \ $n\sum_{i=1}^n X_{i-1}^2 - \left(\sum_{i=1}^n X_{i-1}\right)^2\ne 0$. \ Since the matrix \[ \begin{bmatrix} 2\sum_{i=1}^n X_{i-1}^2 & 2\sum_{i=1}^n X_{i-1} \\ 2\sum_{i=1}^n X_{i-1} & 2n \end{bmatrix} \] consisting of the second order partial derivatives of the function \ $\RR^2\ni(\gamma,\delta)\mapsto \sum_{i=1}^n (X_i - \gamma X_{i-1} - \delta)^2$ \ is positive definite provided that \ $n\sum_{i=1}^n X_{i-1}^2 - \left(\sum_{i=1}^n X_{i-1}\right)^2>0$, \ we have \ $(\widehat\gamma_n^{\mathrm{CLSE}},\widehat \delta_n^{\mathrm{CLSE}})$ \ is indeed the solution of the extremum problem provided that \ $n\sum_{i=1}^n X_{i-1}^2 - \left(\sum_{i=1}^n X_{i-1}\right)^2>0$. \begin{Rem}\label{Rem5} Using the definition of CLSE of \ $(\theta,m)$ \ we give a mathematical motivation of the definition of the LSE \ $\widetilde\theta_n$ \ of \ $\theta$ \ introduced in Section \ref{section_LSE_theta}. Note that if \ $\theta=0$, \ then \[ X_i - \EE( X_i \mid \cF^{(Y,X)}_{i-1}) = X_i - X_{i-1} - m, \quad i=1,\ldots,n. \] If \ $\theta\ne 0$, \ then, by Taylor's theorem, \ $1-\ee^{-\theta} = \ee^{-\tau\theta}\theta$ \ with some \ $\tau=\tau(\theta)\in[0,1]$, \ and hence \begin{align*} X_i - \EE( X_i \mid \cF^{(Y,X)}_{i-1}) &= X_i - \ee^{-\theta} X_{i-1} - m \int_0^1\ee^{-\theta v}\,\dd v\\ &= X_i - X_{i-1} + \ee^{-\tau\theta}\theta X_{i-1} - m\ee^{-\tau\theta} \end{align*} for \ $i=1,\ldots,n-1$. \ Hence for small values of \ $\theta$ \ one can approximate \ $X_i - \EE( X_i \mid \cF^{(Y,X)}_{i-1})$ \ by \ $X_i - X_{i-1} + \theta X_{i-1} - m = X_i - X_{i-1} - (m-\theta X_{i-1})$, $i=1,\ldots,n$. \ Based on this, for small values of \ $\theta$, \ in the definition of the LSE of \ $\theta$, \ the sum \ $\sum_{i=1}^n (X_i - X_{i-1} - (m-\theta X_{i-1}))^2$ \ can be considered as an approximation of the sum \ $\sum_{i=1}^n ( X_i - \EE( X_i \mid \cF^{(Y,X)}_{i-1}) )^2$ \ in the definition of the CLSE of \ $(\theta,m)$. \proofend \end{Rem} \begin{Thm}\label{Thm4} Let us assume that Condition (C) holds. Then \begin{align}\label{help28} \PP\left(n\sum_{i=1}^n X_{i-1}^2 - \left(\sum_{i=1}^n X_{i-1}\right)^2>0\right)=1 \qquad \text{for all \ $n\geq 2$,} \end{align} and there exists a unique CLSE \ $(\widehat\theta_n^{\mathrm{CLSE}},\widehat m_n^{\mathrm{CLSE}})$ \ which has the form given in \eqref{help29}. Further, \begin{align}\label{help30} n\widehat\theta_n^{\mathrm{CLSE}} \distr - \frac{\int_0^1 \cX_t\,\dd \cX_t - \cX_1\int_0^1 \cX_t\,\dd t} {\int_0^1 \cX_t^2\,\dd t - \left(\int_0^1 \cX_t\,\dd t\right)^2} \qquad \text{as \ $n\to\infty$,} \end{align} and \begin{align}\label{help31} \widehat m_n^{\mathrm{CLSE}} \distr \frac{\cX_1\int_0^1 \cX_t^2\,\dd t - \int_0^1 \cX_t\,\dd t \int_0^1 \cX_t\,\dd \cX_t} {\int_0^1 \cX_t^2\,\dd t - \left(\int_0^1 \cX_t\,\dd t\right)^2} \qquad \text{as \ $n\to\infty$,} \end{align} where \ $(\cX_t)_{t\in\RR_+}$ \ is the second coordinate of a two-dimensional affine process\ $(\cY_t,\cX_t)_{t\in\RR_+}$ \ given by the unique strong solution of the SDE \begin{align*} \begin{cases} \dd\cY_t = a\,\dd t + \sqrt{\cY_t}\,\dd \cW_t,\\[2mm] \dd\cX_t = m\,\dd t + \sqrt{\cY_t}\,\dd \cB_t, \end{cases} \qquad t\in\RR_+, \end{align*} with initial value \ $(\cY_0,\cX_0)=(0,0)$, \ where \ $(\cW_t)_{t\in\RR_+}$ \ and \ $(\cB_t)_{t\in\RR_+}$ \ are independent standard Wiener processes. \end{Thm} \noindent{\bf Proof.} By the proof of Theorem \ref{Thm3}, we have \eqref{help28}. By \eqref{help43} and \eqref{help44}, for all \ $n\geq 2$ \ we have \begin{align*} \widehat\gamma_n^{\mathrm{CLSE}} - 1 = \frac{n\sum_{i=1}^n (X_i - X_{i-1})X_{i-1} - X_n\sum_{i=1}^n X_{i-1}} {n\sum_{i=1}^n X_{i-1}^2 - \left( \sum_{i=1}^n X_{i-1} \right)^2}, \end{align*} and \begin{align*} \widehat\delta_n^{\mathrm{CLSE}} = \frac{X_n\sum_{i=1}^n X_{i-1}^2 - \sum_{i=1}^n X_{i-1}\sum_{i=1}^n (X_i - X_{i-1})X_{i-1}} {n\sum_{i=1}^n X_{i-1}^2 - \left( \sum_{i=1}^n X_{i-1} \right)^2}. \end{align*} Using \eqref{help53} and \eqref{help10}, the continuous mapping theorem, by the same technique as in the proof of Theorem \ref{Thm3}, we get \begin{align}\label{help32} n(\widehat\gamma_n^{\mathrm{CLSE}} - 1) \distr \frac{\int_0^1 \cX_t\,\dd \cX_t - \cX_1\int_0^1 \cX_t\,\dd t} {\int_0^1 \cX_t^2\,\dd t - \left(\int_0^1 \cX_t\,\dd t\right)^2} \qquad \text{as \ $n\to\infty$,} \end{align} and \begin{align}\label{help33} \widehat \delta_n^{\mathrm{CLSE}} \distr \frac{\cX_1\int_0^1 \cX_t^2\,\dd t - \int_0^1 \cX_t\,\dd t \int_0^1 \cX_t\,\dd \cX_t} {\int_0^1 \cX_t^2\,\dd t - \left(\int_0^1 \cX_t\,\dd t\right)^2} \qquad \text{as \ $n\to\infty$.} \end{align} By Slutsky's lemma, we also have \ $\widehat\gamma_n^{\mathrm{CLSE}} \stoch 1$ \ as \ $n\to\infty$. \ Hence, by Taylor's theorem using also that \ $\widehat\gamma_n^{\mathrm{CLSE}}>0$, $n\in\NN$ \ (due to its definition given in \eqref{help29}), we have \begin{align}\label{help34} \widehat\theta_n^{\mathrm{CLSE}} = - \log(\widehat\gamma_n^{\mathrm{CLSE}}) = - \log(\widehat\gamma_n^{\mathrm{CLSE}}) - \log(1) = -\frac{1}{\xi_n}(\widehat\gamma_n^{\mathrm{CLSE}} - 1), \end{align} where \ $\xi_n$ \ is in the interval with endpoints \ $1$ \ and \ $\widehat\gamma_n^{\mathrm{CLSE}}$. \ Since \ $\widehat\gamma_n^{\mathrm{CLSE}}\stoch 1$ \ as \ $n\to\infty$, \ we have \ $\xi_n\stoch 1$ \ as \ $n\to\infty$, \ and hence using the decomposition \begin{align*} n \widehat\theta_n^{\mathrm{CLSE}} = - \frac{1}{\xi_n}n(\widehat\gamma_n^{\mathrm{CLSE}} - 1), \qquad n\in\NN, \end{align*} Slutsky's lemma and \eqref{help32}, we get \eqref{help30}. Next we turn to prove \eqref{help31}. For this, by \eqref{help29}, \eqref{help33} and by Slutsky's lemma, it is enough to check that \[ \int_0^1 \ee^{-\widehat\theta_n^{\mathrm{CLSE}}v}\,\dd v \stoch 1 \qquad \text{as \ $n\to\infty$.} \] Since \begin{align*} \int_0^1 \ee^{-\widehat\theta_n^{\mathrm{CLSE}}v}\,\dd v = \begin{cases} \frac{1-\ee^{-\widehat\theta_n^{\mathrm{CLSE}}}} {\widehat\theta_n^{\mathrm{CLSE}}} & \text{if \ $\widehat \theta_n^{\mathrm{CLSE}}\ne 0 $,}\\ 1 & \text{if \ $\widehat \theta_n^{\mathrm {CLSE}}=0$,} \end{cases} \end{align*} by \eqref{help34}, for all \ $\varepsilon>0$ \ we have \begin{align*} \PP\left( \left\vert \int_0^1 \ee^{-\widehat\theta_n^{\mathrm {CLSE}}v}\,\dd v -1 \right\vert \geq \varepsilon\right) &= \PP\left( \left\vert \frac{1-\widehat\gamma_n^{\mathrm {CLSE}}}{\widehat\theta_n^{\mathrm {CLSE}}} -1 \right\vert \geq \varepsilon \,\Big\vert\, \widehat\theta_n^{\mathrm {CLSE}}\ne 0 \right) \PP(\widehat\theta_n^{\mathrm {CLSE}}\ne 0)\\ &\phantom{=\;} + \PP\left( \vert 1-1 \vert \geq \varepsilon \,\big \vert\, \widehat\theta_n^{\mathrm {CLSE}}= 0 \right) \PP(\widehat\theta_n^{\mathrm {CLSE}}= 0)\\ & = \PP\left( \vert \xi_n-1 \vert \geq \varepsilon \,\big \vert\, \widehat\theta_n^{\mathrm {CLSE}}\ne 0 \right) \PP(\widehat\theta_n^{\mathrm {CLSE}}\ne 0) \stoch 0, \end{align*} since \ $\xi_n\stoch 0$ \ as \ $n\to\infty$. \proofend \begin{Rem} (i) We do not consider the CLSE of \ $\theta$ \ supposing that \ $m$ \ is known since the corresponding extremum problem is rather complicated, and from statistical point of view it has less importance. \noindent(ii) Under the Condition (C), by Theorem \ref{Thm4} and Slutsky's lemma, we get \ $\widehat\theta_n^{\mathrm{CLSE}}$ \ converges stochastically to the parameter \ $\theta=0$ \ as \ $n\to\infty$, \ and one can show that \ $\widehat m_n^{\mathrm{CLSE}}$ \ does not converge stochastically to the parameter \ $m$ \ as \ $n\to\infty$, \ see Appendix \ref{AppendixB}. \proofend \end{Rem} \medskip
\section{Introduction} In this work we consider an interesting bosonization of ${\mathcal N}$=1 Liouville field theory that was proposed recently in \cite{Belavin:2011sw}. ${\mathcal N}$=1 Liouville field theory contains one fermionic field $\psi$ in addition to the Liouville field $\varphi$. These fields are coupled through the standard interaction term. For bosonization we need to add another free fermion $\eta$. The product theory appears naturally in several applications of ${\mathcal N}$=1 Liouville field theory. In particular, it has been used in \cite{Hikida:2007sz} and \cite{Creutzig:2010zp} to compute various structure constants of the OSP(1$|$2) WZW model. More recently, it was considered in the context of the AGT correspondence \cite{Alday:2009aq} between supersymmetric 4D gauge theories and 2D conformal field theory \cite{Belavin:2011pp, Nishioka:2011jk, Bonelli:2011jx, Bonelli:2011kv, Belavin:2011sw}. In the bosonization, the two fermionic fields $\psi$ and $\eta$ are replaced by a single boson $Y$. What Belavin et al. proposed was that the two bosonic fields $\varphi$ and $Y$ can me mapped to a new set of bosonic fields, $X$ and $\hat X$, where $X$ is an ordinary (non-suspersymmetric) Liouville field and $\hat X$ an imaginary cousin. The latter may be thought of as a Liouville field which takes values in imaginary numbers. Because of its internal structure, we shall often refer to the fully bosonic model as {\em double Liouville theory} and to the factor associated with the field $\hat X$ as imaginary Liouville theory. Imaginary Liouville theory is far from being an established model of 2-dimensional conformal field theory. In fact, there exist several different proposals for its structure constants but consistency (crossing symmetry) has never been established (see discussion in section 3). It is remarkable that one version of imaginary Liouville theory now appears through the bosonization of a consistent local conformal field theory. The relation between ${\mathcal N}$=1 and double Liouville theory has a suggestive ancestor in rational conformal field theory. In that context, double Liouville theory gets replaced by a product of two minimal models and ${\mathcal N}$=1 Liouville theory by its rational counterpart. We can give a highly suggestive argument for their relation if we represent both models as coset conformal field theories. It is well known that ordinary minimal models arise through the cosets $$ {\text{MM}}_k = (SU(2)_{k} \times SU(2)_1) / SU(2)_{k+1} $$ where $ k = 1, 2, \dots$. This family of rational models includes the Ising model MM$_1$ for a single fermion $\eta$ when $k=1$. Similarly, ${\mathcal N}$=1 supersymmetric minimal models are obtained from the coset $$ {\text{SMM}}_k = (SU(2)_k \times SU(2)_2)/SU(2)_{k+2}\ . $$ If we allow ourselves to extend and reduce both numerator and denominator by the required additional factors we can easily see that \begin{equation} \label{mainrat} {\text{SMM}}_{k-1} \times {\text{MM}}_1 \sim {\text{MM}}_k \times {\text{MM}}_{k-1} \ . \end{equation} Similar relations between 'generalized minimal models' and Virasoro minimal models were first discussed in \cite{Crnkovic:1989gy}, \cite{Crnkovic:1989ug} and later (it seems independently) by \cite{Lashkevich:1992sb},\cite{Lashkevich:1993fb}. More recently, results for the 4D gauge theories \cite{Bonelli:2011kv} inspired Wyllard \cite{Wyllard:2011mn} to propose an extension to cosets of the type $(SU(N)_\kappa \times SU(N)_p )/SU(N)_{\kappa+p}$ where $\kappa$ is a free parameter. Soon after this paper had appeared, the case of $N=2, p=2$ was considered in more detail by Belavin et al. \cite{Belavin:2011sw}. Let us now describe the content of this work in more detail. We shall begin with a brief review of Liouville field theory and its ${\mathcal N}$=1 supersymmetric version in the next section. Both theories were solved long ago, see section 2 for references to the original literature. Then we turn to imaginary Liouville theory. As mentioned before, this model is very poorly understood. After a few historical comments we shall describe the 3-point functions that were proposed by Zamolodchikov in \cite{Zamolodchikov:2005fy}. Our new results are formulated and analyzed in section 4. There we shall spell out a precise relation between an infinite tower of fields in ${\mathcal N}$=1 Liouville field theory and double Liouville theory. This relation will be checked through extensive comparison of 3-point functions on both sides of the correspondence. Applications and extensions of our results are sketched in the concluding section. \section{Review of Liouville field theory} \def\mathcal{B}{\mathcal{B}} In this section we simply review some basic facts about Liouville field theory and its ${\mathcal N}$=1 supersymmetric cousin. Most importantly, we shall discuss the spectrum of primary fields along with their 2- and 3-point functions. For a more details see the reviews \cite{Teschner:2001rv,Schomerus:2005aq,Nakayama:2004vk}. \subsection{Bosonic Liouville field theory} \def\tilde \alpha{\tilde \alpha} Liouville field theory involves a single scalar field with an exponential interaction term. On a 2-dimensional world-sheet with metric ${\gamma}^{ab}$ and curvature $R$, the action of Liouville theory takes the form \begin{equation} \label{actLiouv} S_L[X] \ = \ \frac{1}{4\pi}\int_\Sigma d^2 \sigma \sqrt {\gamma} \left( {\gamma}^{ab} \partial_a X \partial_b X + R Q X + 4 \pi \mu_L e^{2bX} \right) \end{equation} where $\mu_L$ and $b$ are two (real) parameters of the model. The second term in this action describes the background charge of a linear dilaton. The value of the constant $Q$ must be adjusted to the choice of $b$ in order for $S_L$ to define a conformal quantum field theory. We shall state the relation in a moment. Liouville theory should be considered as a marginal deformation of the free linear dilaton theory. The Virasoro field of a linear dilaton theory is given by the familiar expression $$ T(z) \ = \ - (\partial X)^2 + Q \partial^2 X \ \ . $$ The modes of this field form a Virasoro algebra with central charge $c_L = 1 + 6 Q^2$. Furthermore, the usual closed string vertex operators \begin{equation} V_\alpha(z) \ =\ :\exp 2\alpha X(z,\bar z): \ \ \ \ \mbox{ have } \ \ h_\alpha \, = \, \alpha (Q-\alpha)\, = \, \bar \Lcd_\alpha\ \ . \label{Lcd} \end{equation} Here and in the following we shall not explicitly display the dependence of our vertex operators on the complex conjugate $\bar z$ of the world-sheet coordinate $z$. Note the conformal weights $h,\bar \Lcd$ are real if $\alpha$ is of the form $\alpha = Q/2 + iP$. In order for the exponential potential in the Liouville action to be marginal, i.e.\ $(h_b,\bar \Lcd_b) = (1,1)$, we must now also adjust the parameter $Q$ to the choice of $b$ in such a way that $$ Q = b + b^{-1} \ \ . $$ Weyl invariance of the classical action $S_L$ leads to the relation $Q_{c} = b^{-1}$ and the additional shift by $b$ may be considered as a quantum correction of the classical relation. The extra term, which certainly becomes small in the semi-classical limit $b \rightarrow 0$, renders $Q = Q_c + b$ (and hence the central charge) invariant under the replacement $ b \rightarrow b^{-1}$. The solution of Liouville field theory is completely described by the 2- and 3-point functions of the model. The vertex operators $V_\alpha$ are introduced such that their 2-point function is canonically normalized, i.e. \begin{equation} \langle V_{\alpha_2}(z_2) V_{\alpha_1} (z_1) \rangle = |z_{12}|^{-4h_{\alpha_1}} 2 \pi \left( \delta(\alpha_1+\alpha_2-Q) + D_L(\alpha_1) \delta(\alpha_2-\alpha_1)\right) \end{equation} where \begin{equation} \label{L2pt} D_L(\alpha) = \left(\pi \mu_L \gamma(b^2) \right)^{(Q-2\alpha) \over b} { \gamma(2 \alpha b - b^2) \over b^2 \, \gamma(2 - 2\alpha b^{-1}+ b^{-2})} \end{equation} Here and throughout the main text we use $\gamma(x) = \Gamma (x)/ \Gamma(1-x)$. In order to spell out the 3-point functions we need to introduce Barnes' double $\Gamma$-function $\Gamma_b(y)$. It may be defined through the following integral representation, \begin{equation}\label{BGamma} \ln \Gamma_b(y) \ = \ \int_0^\infty \frac{d\tau}{\tau} \left[ \frac{e^{-y\tau} - e^{- Q \tau/2}} {(1-e^{-b \tau}) (1- e^{-\tau/b})} - \frac{\left(\frac{Q}{2} - y\right)^2}{2} e^{-\tau} - \frac{\frac{Q}{2} -y }{\tau}\right] \end{equation} for all $b \in \mathbb{R}$. The integral exists when $0 < {\rm Re}(y)$ and it defines an analytic function which may be extended onto the entire complex $y$-plane. Under shifts by $b^{\pm 1}$, the function $\Gamma_b$ behaves according to \begin{equation}\label{BGshift} \Gamma_b(y+b) \ = \ \sqrt{2\pi} \, \frac{b^{by-\frac12}}{\Gamma(by)}\, \Gamma_b(y) \ \ , \ \ \Gamma_b(y+b^{-1}) \ = \ \sqrt{2\pi} \, \frac{b^{-\frac{y}{b}+\frac12}} {\Gamma(b^{-1}y)}\, \Gamma_b(y) \ \ . \end{equation} These shift equations let $\Gamma_b$ appear as an interesting generalization of the usual $\Gamma$ function which may also be characterized through its behavior under shifts of the argument. But in contrast to the ordinary $\Gamma$ function, Barnes' double $\Gamma$ function satisfies two such equations which are independent if $b$ is not rational. We furthermore deduce from eqs.\ (\ref{BGshift}) that $\Gamma_b$ has poles at \begin{equation} \label{poles} y_{n,m} \ = \ - n b - m b^{-1} \ \ \ \mbox{ for } \ \ \ n,m \ = \ 0,1,2, \dots \ \ . \end{equation} From Branes' double Gamma function one may construct the following basic building block of the 3-point function, \begin{equation} \label{ZY} \Upsilon_b (\alpha) \ := \ \Gamma_2(\alpha|b,b^{-1})^{-1}\, \Gamma_2(Q-\alpha|b,b^{-1})^{-1} \ \ . \end{equation} The properties of the double $\Gamma$-function imply that $\Upsilon$ possesses the following integral representation \begin{equation}\label{ZYpsilon} \ln \Upsilon_b(y) \ = \ \int_0^\infty \frac{dt}{t} \left[ \left(\frac{Q}{2}-y\right)^2 e^{-t} - \frac{\sinh^2\left(\frac{Q}{2} -y\right) \frac{t}{2}} {\sinh\frac{bt}{2}\, \sinh \frac{t}{2b}}\right]\ \ . \end{equation} Moreover, we deduce from the two shift properties (\ref{BGshift}) of the double $\Gamma$-function that \begin{equation}\label{ZYshift} \Upsilon_b(y+b) \ = \ \gamma(by) \, b^{1-2by}\, \Upsilon_b(y) \ \ , \ \ \Upsilon_b(y+b^{-1}) \ = \ \gamma(b^{-1}y) \, b^{-1+2b^{-1}y}\, \Upsilon_b(y) \ \ . \end{equation} Note that the second equation can be obtained from the first with the help of the self-duality property $\Upsilon_b(y) = \Upsilon_{b^{-1}}(y)$. \smallskip After this preparation it is easy to spell out the 3-point function of primary fields in Liouville field theory \cite{Dorn:1994xn,Zamolodchikov:1995aa}, \begin{equation}\label{Cc>} \langle V_{\alpha_3}(z_3) V_{\alpha_2}(z_2) V_{\alpha_1} (z_1) \rangle = \frac{C_L(\alpha_3,\alpha_2,\alpha_1|b)} {|z_{12}|^{2h_{12}} |z_{13}|^{2h_{13}} |z_{23}|^{2h_{23}}} \end{equation} with $h_{12} = h_{\alpha_1}+ h_{\alpha_2}-h_{\alpha_3}$ etc. and coupling constants $C_L$ of the form \begin{equation}\label{Lstr} C_L(\alpha_3,\alpha_2,\alpha_1|b) = \left[\pi \mu_L \gamma(b^2) b^{2-2b^2} \right]^{\frac{Q-\talpha}{b}} \!\! \frac{\Upsilon^0_b \, \Upsilon_{b}\left( 2 \alpha_1\right) \Upsilon_{b}\left( 2 \alpha_2\right) \Upsilon_{b}\left( 2 \alpha_3\right) }{ \Upsilon_{b}\left( \talpha_{123} - Q\right)\, \Upsilon_{b}\left( \talpha_{12}\right)\, \Upsilon_{b}\left( \talpha_{13}\right)\, \Upsilon_{b}\left( \talpha_{23}\right) }\ . \end{equation} Here and in the following, the contant $\Upsilon^0_b$ is given by $\Upsilon^0_b = \Upsilon'_{b}\left( 0\right)$. Furthermore, the parameters $\talpha_{123}$ and $\talpha_{ij}$ are certain linear combinations of $\alpha_j$, $$ \talpha_{123} = \alpha_1+\alpha_2+\alpha_3 \ , \ \talpha_{12} = \alpha_1 + \alpha_2 - \alpha_3 \quad \mbox{etc.}$$ The solution (\ref{Lstr}) was first proposed by H.\ Dorn and H.J.\ Otto \cite{Dorn:1994xn} and by A.\ and Al.\ Zamolodchikov \cite{Zamolodchikov:1995aa}, based on extensive earlier work by many authors (see e.g.\ the reviews \cite{Seiberg:1990eb,Teschner:2001rv, Schomerus:2005aq} for references). Full crossing symmetry of the conjectured 3-point function was established much later in two steps by Ponsot and Teschner \cite{Ponsot:1999uf} and by Teschner \cite{Teschner:2001rv,Teschner:2003en}. The proof of consistency of the DOZZ structure constants for Liouville field theory was completed recently by establishing modular invariance of 1-point functions on a torus \cite{Hadasz:2009sw}. \subsection{${\mathcal N}$=1 Liouville field theory} ${\mathcal N}$=1 supersymmetric Liouvilel field involves one real superfield that contains a real bosonic scalar $\varphi$, the two components $\psi$ and $\bar \psi$ of a Majorana fermion and an auxiliary field $F$. After integrating out the latter and fixing the world-sheet metric, the action of ${\mathcal N}$=1 super Liouville field theory takes the form \begin{align} S_{SL}[\varphi,\psi] \ = \ \frac{1}{2\pi} \int d^2 z \left[ \partial \varphi \bar \partial \varphi + \psi \bar \partial \psi + \bar \psi \partial \bar \psi \right] + 2 i \mu b^2 \int d^2 z \psi \bar \psi e^{ b \varphi} ~, \end{align} The background charge for the boson $\varphi$ is related to the parameter $b$ by $Q = b + 1/b$. As in the case of bosonic Liouville field theory, the supersymmetric cousin is obtained by perturbing a free field theory, namely the product of a linear dilaton with a 2-dimensional Ising model. The spectrum of the Ising model contains six conformal blocks including the identity field, the two components $\psi$ and $\bar \psi$ of the fermion and the energy density $\psi \bar \psi$, which are all part of the Neveu-Schwarz (NS) sector. In addition, there are two blocks in the Ramond (R) sector. These are generated from the spin field $\varsigma^+ = \sigma$ and the so-called disorder field $\varsigma^- = \mu$. After multiplication with the linear dilator, the model contains an ${\mathcal N}$=1 super-conformal symmetry with central charge $c_{SL} = \frac{3}{2}(1 + 2 Q^2)$. The holomorphic half of this symmetry is generated by modes of the following fields \begin{equation} T(z) = - \frac12\left((\partial \varphi)^2 - Q \partial^2 \varphi + \psi \partial \psi\right) \quad , \quad G(z) = -i( \psi \partial \varphi - Q \partial \psi ) \ . \label{N1SCA} \end{equation} Anti-holomorphic fields can be constructed similarly. The interacting theory has been solved soon after the DOZZ proposal had been put out, see \cite{Rashkov:1996jx,Poghosian:1996dw}. Vertex operators in the NS sector are super-descendents of \begin{equation} \phi_\alpha (z) = : \exp \alpha \varphi(z,\bar z) : \quad \mbox{ with } \quad \Delta_\alpha = \alpha(Q-\alpha)/2 = \bar \Delta_\alpha \label{NSvert} \end{equation} The 2-point function of these NS primary fields takes the form \begin{align}\label{L2ptNS} \langle \phi_{\alpha_2} (z_2) \phi_{\alpha_1} (z_1 ) \rangle \ =\ |z_{12}|^{-4 \Delta_{\alpha_1}} 2 \pi \left[ \delta (\alpha_1 + \alpha_2 - Q) + \delta (\alpha_2 - \alpha_1 ) D_{NS}(\alpha_1) \right] ~, \end{align} with \begin{align} D_{NS} (\alpha) \ = \ - \left(\mu \pi \gamma (\tfrac{bQ}{2}) \right)^{\frac{Q-2\alpha}{b}} \frac{\Gamma ( b (\alpha-\frac{Q}{2}) ) \Gamma ( \frac{1}{b} (\alpha-\frac{Q}{2}) ) } {\Gamma (- b (\alpha-\frac{Q}{2}) ) \Gamma ( - \frac{1}{b} (\alpha-\frac{Q}{2}) ) } ~. \end{align} Whereas the first term in eq.\ \eqref{L2ptNS} is fixed by normalization, the second term involving $D_{NS}$ contains dynamical information on the phase shift of tachyonic modes upon reflection off the Liouville wall. \smallskip To spell out the 3-point functions of the model we need to build two new special functions from the $\Upsilon$-function we introduced in the previous subsection, see eq.\ \eqref{ZYpsilon}. These are given by \begin{align} \Upsilon^{\text{NS}}_b (x) &= \Upsilon_b ( \tfrac{x}{2}) \Upsilon_b (\tfrac{x+Q}{2}) ~, &\Upsilon^{\text{R}}_b (x) &= \Upsilon_b (\tfrac{x+b}{2}) \Upsilon_b (\tfrac{x+b^{-1}}{2}) ~. \end{align} Properties of these new functions can easily be derived from the properties of $\Upsilon_b$ we listed above. In particular, we note that the functions $\Upsilon^{\text NS}_b$ and $\Upsilon^{\text R}_b$ possess the following behavior under shifts of their argument, \begin{align} \Upsilon^{\text{NS}}_b (x+b) &= b^{-b x} \gamma ( \tfrac12 + \tfrac{bx}{2}) \Upsilon^{\text{R}}_b(x) ~, &\Upsilon^{\text{R}}_b (x+b) &= b^{1-b x} \gamma (\tfrac{bx}{2}) \Upsilon^{\text{NS}}_b (x)~, \\[2mm] \Upsilon^{\text{NS}}_b (x+ \tfrac{1}{b}) &= b^{\frac{x}{b}} \gamma (\tfrac12 + \tfrac{x}{2b}) \Upsilon^{\text{R}}_b (x)~, &\Upsilon^{\text{R}}_b (x+ \tfrac{1}{b}) &= b^{-1+ \frac{x}{b} } \gamma (\tfrac{x}{2b}) \Upsilon^{\text{NS}}_b (x) ~. \label{uprel} \end{align} The functions $\Upsilon_b^{\rm NS}, \Upsilon^{\text{R}}_b$ suffice to state the 3-point structure constants of the NS sector, \begin{equation}\label{NS3p} \langle \phi_{\alpha_3}(z_3) \phi_{\alpha_2}(z_2) \phi_{\alpha_1} (z_1) \rangle = \frac{C_{NS}(\alpha_3,\alpha_2,\alpha_1|b)} {|z_{12}|^{2h_{12}} |z_{13}|^{2h_{13}} |z_{23}|^{2h_{23}}} \end{equation} \begin{equation} \label{NS3p2} \langle \phi_{\alpha_3}(z_3) \tilde \phi_{\alpha_2}(z_2) \phi_{\alpha_1}(z_1) \rangle = \frac{\tilde C_{NS}(\alpha_3,\alpha_2,\alpha_1|b)} {|z_{12}|^{2h_{12}+1} |z_{13}|^{2h_{13}-1} |z_{23}|^{2h_{23}+1}} \end{equation} where $\tilde \phi_\alpha = \{ G_{-\frac12},[ \bar G_{-\frac12} ,\phi_\alpha]\}$, and \begin{eqnarray} C_{NS}(\alpha_3,\alpha_2,\alpha_1|b)&=&\frac12 \left[\frac{\pi \mu}{2b^{b^2-1}} \gamma\left(\frac{Qb}{2}\right) \right]^{Q-\talpha_{123} \over b}\!\!\!\!\!\! \frac{\Upsilon^0_b \,\Upsilon_b^{\rm NS}(2\alpha_1) \Upsilon_b^{\rm NS}(2\alpha_2) \Upsilon_b^{\rm NS}(2\alpha_3)}{ \Upsilon_b^{\rm NS}(\talpha_{123}-Q) \Upsilon_b^{\rm NS}(\talpha_{12})\Upsilon_b^{\rm NS}(\talpha_{23})\Upsilon_b^{\rm NS}(\talpha_{13})}\nonumber\\[2mm] & & \label{SLstr} \\[-2mm] \tilde C_{NS}(\alpha_3, \alpha_2, \alpha_1|b)&=& i \left[ \frac{\pi \mu}{2b^{b^2-1}} \gamma\left(\frac{Qb}{2}\right)\right]^{Q-\talpha_{123} \over b} \!\!\!\!\!\! \frac{\Upsilon^0_b \,\Upsilon_b^{\rm NS}(2\alpha_1) \Upsilon_b^{\rm NS}(2\alpha_2) \Upsilon_b^{\rm NS}(2\alpha_3)}{ \Upsilon_b^{\rm R}(\talpha_{123}-Q) \Upsilon_b^{\rm R}(\talpha_{12})\Upsilon_b^{\rm R}(\talpha_{23})\Upsilon_b^{\rm R}(\talpha_{13})} \nonumber \end{eqnarray} Any 3-point function of descendent fields can be written in terms of the correlator (\ref{NS3p}) or (\ref{NS3p2}) and 3-point blocks which are completely determined by the super-conformal Ward identities (see e.g. \cite{Hadasz:2006qb}). Let us now turn to the R sector of the model. As we recalled before, the 2-dimensional Ising model possesses two local fields of conformal weight $\Delta = 1/16 = \bar \Delta$ which we denoted by $\varsigma^+ = \sigma$ and $\varsigma^- = \mu $, see chapter 12 of \cite{DiFrancesco:1997nk} for more details. Using these spin fields, we can define the following two vertex operators in the R sector of ${\mathcal N}$=1 Liouville theory \begin{align} \Sigma^{\pm}_{\alpha}(z) &=\ \varsigma^{\pm}(z,\bar z) \ :e^{ \alpha \varphi(z,\bar z)}: \quad \mbox{with} \quad \Delta_\alpha^{\rm R} = \frac12\alpha(Q-\alpha) + \frac{1}{16} = \bar \Delta^{\rm R}_\alpha \ . \label{spin} \end{align} Our conventions are the same as in \cite{Poghosian:1996dw, Hadasz:2008dt} and they imply \begin{eqnarray}\label{Rfields} G_0 \Sigma^\pm_{\alpha}(z) = i \beta e^{\mp i \frac{\pi}{4} } \Sigma^\mp_{\alpha}(z) , \quad \bar G_0 \Sigma^\pm_{\alpha}(z) = -i \beta e^{\pm i \frac{\pi}{4} } \Sigma^\mp_{\alpha}(z) , \qquad \beta = \frac{1}{\sqrt2} \left(\frac{Q}{2} - \alpha \right)\, . \end{eqnarray} The 2-point functions of the vertex operators $\Sigma^{\epsilon}_{\alpha}$ possess the following form \begin{align} &\langle \Sigma^{\pm}_{\alpha_2} (z_2) \Sigma^{\pm}_{\alpha_1} (z_1) \rangle \ = \ |z_{12}|^{- 4 \Delta_{\alpha_1} - \frac14} 2 \pi \left[\delta (\alpha_1 + \alpha_2 - Q) \pm \delta (\alpha_2 - \alpha_1 ) D_R (\alpha_1) \right] ~ \label{L2ptR} \end{align} with a reflection coefficient given by \begin{align} D_R (\alpha )\ = \ \left(\mu \pi \gamma (\tfrac{bQ}{2}) \right)^{\frac{Q-2\alpha}{b}} \frac{\Gamma (\frac12 + b (\alpha-\frac{Q}{2}) ) \Gamma (\frac12 + \frac{1}{b} (\alpha-\frac{Q}{2}) ) } {\Gamma (\frac12 - b (\alpha-\frac{Q}{2}) ) \Gamma (\frac12 - \frac{1}{b} (\alpha-\frac{Q}{2}) ) } ~. \label{refR} \end{align} Let us also provide explicit expressions for the 3-point functions involving two RR fields. These were determined in \cite{Rashkov:1996jx, Poghosian:1996dw,Fukuda:2002bv} and we shall simply quote the results along with all the necessary notations, \begin{align}\label{SLstrR} \langle \phi_{\alpha_3} (z_3) \Sigma_{\alpha_2}^{\pm} (z_2) \Sigma_{\alpha_1}^{\pm} (z_1) \rangle \ = \ \frac{C^{\pm}_R (\alpha_3 ; \alpha_2 , \alpha_1|b)}{|z_{12}|^{2\Delta_{12}+\frac14} |z_{23}|^{2\Delta_{23}} |z_{13}|^{2\Delta_{13}}} ~. \end{align} The structure constants $C^{\pm}_R$ are constructed from the special functions $\Upsilon^{\rm NS}$ and $\Upsilon^{\rm R}$ as follows, \begin{eqnarray}\label{N1LstrR} \nonumber C^{\pm}_R (\alpha_3; \alpha_2, \alpha_1|b ) &=&\frac12 \left[ \frac{\mu \pi}{2} \gamma (\tfrac{bQ}{2} ) b^{1-b^2} \right]^{\frac{Q- \talpha_{123}}{b}} \frac{\Upsilon^{0}_b \, \Upsilon^{\text{R}}_b(2 \alpha_1) \Upsilon^{\text{R} }_b (2 \alpha_2 ) \Upsilon^{\text{NS}}_b (2 \alpha_3) } { \Upsilon^{\text{R}}_b( \talpha_{123} - Q) \Upsilon^{\text{R} }_b (\talpha_{12}) \Upsilon^{\text{NS}}_b (\talpha_{23}) \Upsilon^{\text{NS}}_b (\talpha_{13}) } \\[2pt] \\[-2pt] \nonumber &\pm&\!\!\!\frac12 \left[ \frac{\mu\pi}{2} \gamma ( \tfrac{bQ}{2} ) b^{1-b^2} \right]^{\frac{Q- \talpha_{123}}{b}} \frac{\Upsilon ^{0}_b \, \Upsilon^{\text{R}}_b(2 \alpha_1) \Upsilon^{\text{R} }_b (2 \alpha_2 ) \Upsilon^{\text{NS}}_b (2 \alpha_3) } { \Upsilon^{\text{NS}}_b( \talpha_{123} - Q) \Upsilon^{\text{NS} }_b (\talpha_{12}) \Upsilon^{\text{R}}_b (\talpha_{23}) \Upsilon^{\text{R}}_b (\talpha_{13}) } ~. \end{eqnarray} Crossing symmetry of 4-point functions in the NS sector of ${\mathcal N}$=1 Liouville theory with structure constants (\ref{SLstr}) and (\ref{N1LstrR}) was first checked numerically, see \cite{Belavin:2007gz,Belavin:2007eq}, and later proved analytically in \cite{Hadasz:2007wi,Chorazkiewicz:2008es} using braiding and fusion properties of the 4-point blocks. In the case of 4-point functions containing R fields, crossing symmetry of ${\mathcal N}$=1 Liouville theory was verified numerically in \cite{Suchanek:2010kq}. The first step necessary for an analytical proof was presented in \cite{Chorazkiewicz:2011zd} where braiding properties of the 4-point blocks were derived. \section{Imaginary Liouville theory} Before we can state the main results of this work, we need one more ingredient, namely a version of Liouville field theory with central charge $c \leq 1$. In contrast to the models we described in the previous section, the status of the theory we are about to discuss is less clear. In particular, the issue of crossing symmetry has not been settled. We shall begin our exposition with a few historical comments in the first subsection. Then we continue by listing the proposed structure constants without much further discussion. \subsection{Some comments on history} In usual Liouville theory, the parameter $b$ is taken to be real so that the corresponding central charge $c \geq 25$. The explicit expressions for 2- and 3-point functions admit analytic continuation to complex values of $b$ with a non-vanishing real part. Formally, the central charge takes values $1 < c$ in this regime. Purely imaginary values of $b$ have been a subject of several previous studies mostly because such values are relevant for time-like Liouville field theory and tachyon condensation in string theory, see e.g.\ \cite{Gutperle:2003xf,Strominger:2003fn, Schomerus:2003vv,Fredenhagen:2004cj,McElgin:2007ak,Harlow:2011ny, Giribet:2011zx} and further references in the more recent papers. At least for $b=i$, it is possible to define the theory by taking a limit starting with $b = \epsilon + i$. The resulting theory has central charge $c=1$ and it agrees with a certain limit of unitary minimal models. This limit was shown to satisfy crossing symmetry \cite{Runkel:2001ng}. It is likely that similar limits can be taken for other purely imaginary values of $i$. But even if such limits describe consistent local quantum theories, they would at most be defined for a discrete set of $b$-parameters. There is an alternative approach to defining Liouville theory for imaginary $b$, i.e.\ for $c \leq 1$. In order to describe how this works, let us recall a few facts about the usual construction of the 3-point couplings in Liouville field theory. The main idea is to evaluate crossing symmetry for 4-point functions with three physical and one degenerate field insertions. The operator product of a physical with a degenerate field involves a finite set of terms whose coefficients can be computed in free field theory. More precisely, if we take the degenerate field to be $V_{-b^{\pm 1}/2}$, then the 4-point function must satisfy a second order differential equation and hence only two terms can possibly arise on the left hand side of the operator product, e.g.\ \begin{equation} \label{degOPE} V_\alpha(w,\bar w) \, V_{-b/2}(z,\bar{z}) \ = \ \sum_{\pm} \ \frac{c^{\pm}_b(\alpha)}{|z-w|^{h_\pm}} \ V_{\alpha \mp b/2}(z,\bar{z}) \ + \ \dots \end{equation} where $h_\pm = \mp b \alpha + Q (-b/2 \mp b/2)$. A similar expansion for the second degenerate field is obtained by replacing $b \rightarrow b^{-1}$. We can even be more specific about the operator expansions of degenerate fields because the coefficients $c^\pm$ may be determined through a simple free field computation in the linear dilaton background. One finds that \begin{eqnarray} c^-_b(\alpha) & = & - \mu_L \int d^2z \, \langle \, V_{-b/2}(0,0) \, V_{\alpha}(1,1)\, V_{b}(z,\bar{z})\, V_{Q-b/2-\alpha} (\infty,\infty) \, \rangle_{{\rm LD}} \nonumber \\[2mm] & = & - \mu_L \pi \ \frac{{\gamma}(1+b^2) \, {\gamma}(1-2b\alpha)}{{\gamma}(2+b^2 - 2b\alpha)} \ \ \label{cm} \end{eqnarray} see \cite{Schomerus:2005aq} for more details. The result in the second line is obtained using the explicit integral formulas that were derived by Dotsenko and Fateev. The corresponding field is then degenerate and it possesses an operator product consisting of two terms only. {\em Teschner's trick} converts the crossing symmetry condition into a much simpler algebraic condition. Moreover, since we have already computed the coefficients of operator products with degenerate fields, the crossing symmetry equation is in fact linear in the unknown generic 3-point couplings. One component of these conditions for the degenerate field $V_{-b/2}$ reads as follows \begin{eqnarray} \label{spcross} 0 & = & C_L(\alpha_1 + \frac{b}{2},\alpha_3,\alpha_4)\, c^-_b (\alpha_1) {\cal P}^{--}_{+-} + C_L(\alpha_1 - \frac{b}{2},\alpha_3,\alpha_4) \, c^+_b (\alpha_1) {\cal P}^{++}_{+-} \ \ , \\[2mm] \nonumber \mbox{where} & & \ \ {\cal P}^{\pm\pm}_{+-} \ = \ \Fus{\alpha_1\mp b/2,}{\alpha_3 - b/2}{- b/2\, }{\alpha_3}{\ \ \alpha_1\ }{\alpha_4} \ \Fus{\alpha_1\mp b/2,}{\alpha_3 + b/2}{-b/2\, }{\alpha_3}{\ \ \alpha_1\ }{\alpha_4}\ \ . \end{eqnarray} Note that the combination on the right hand side must vanish because in a consistent model, the off-diagonal bulk mode $(\alpha_4-b/2,\alpha_4+b/2)$ does not exist and hence it cannot propagate in the intermediate channel. The required special entries of the Fusing matrix can be expressed through a combination of $\Gamma$ functions. Once the expressions for $c^\pm$ and ${\cal P}$ are inserted (note that they only involve $\Gamma$ functions), the crossing symmetry condition may be written as follows, \begin{equation} \label{3ptse} \frac{C_L(\alpha_1 + b,\alpha_2,\alpha_3)}{C_L(\alpha_1,\alpha_2,\alpha_3)} \ = \ - \frac{{\gamma}(b(2\alpha_1 +b)){\gamma}(2b\alpha_1)}{\pi \mu_L {\gamma}(1+b^2) {\gamma}(b(\talpha_{123}-Q))} \frac{{\gamma}(b(\talpha_{23}-b))}{{\gamma}(b\talpha_{13}) {\gamma}(b\talpha_{12})} \end{equation} with ${\gamma}(x) = \Gamma(x)/ \Gamma(1-x)$, as before. The constraint takes the form of a shift equation that describes how the coupling changes if one of its arguments is shifted by $b$. Using the symmetry $b \leftrightarrow b^{-1}$ we obtain a second shift equation that encodes how the 3-point couplings behave under shifts by $b^{-1}$. For irrational values of $b$, the two shift equations determine the couplings completely, at least if we require that they are analytic in the momenta. The unique solution turns out to be analytic in $b$ as well so that it may be extended to all real values of the parameter $b$. We are now prepared to take a fresh look at the problem of constructing imaginary Liouville theory. While the structure constants \eqref{Lstr} are not analytic in $b$ so that their extension to imaginary $b$ (or $c \leq 1$) may be ill-defined, the coefficients of the shift equation \eqref{3ptse} involve only $\Gamma$ functions so that a continuation to imaginary values of $b$ is straight forward. If we postulate that the 3-point couplings of imaginary Liouville theory are analytic in the parameters $\alpha_i$ and exists for all $c \leq 1$, then there is again a unique solution \cite{Zamolodchikov:2005fy}. We shall describe this solution in the following subsection. \subsection{Zamolodchikov's solution} Imaginary Liouville theory may be thought of as a model whose action is formally given by \begin{equation} \label{actILiouv} S_{\mathcal L}[\hat X] \ = \ \frac{1}{4\pi}\int_\Sigma d^2 \sigma \sqrt {\gamma} \left( - {\gamma}^{ab} \partial_a \hat X \partial_b \hat X + R \hat Q \hat X + 4 \pi \mu_\mathcal{L} e^{-2\hat b\hat X} \right) \end{equation} One can obtain it the usual action of ordinary Liouville theory by the formal replacements $X \rightarrow - i \hat X$, $b \rightarrow -i \hat b$ and $Q \rightarrow i \hat Q$. Vertex operators in this model take the form \begin{equation} {\mathcal V}_{\hat \alpha}(z) = : e^{2 \hat a \hat X(z,\bar z) } \quad \mbox{ with } \quad \hat h_{\hat \alpha} = - \hat \alpha (\hat Q - \hat \alpha)= \hat \bar \Lcd_{\hat \alpha} \ . \label{ILcd} \end{equation} They are obtained from the vertex operators of ordinary Liouville theory if we replace $\alpha$ by $\alpha \rightarrow -\hat \alpha$. For conformal invariance, the parameter $\hat Q$ must be adjusted to the parameter $\hat b$ such that \begin{equation} \hat Q = \hat b^{-1} -\hat b \ \ . \end{equation} In terms of these parameters, the central charge of the Virasoro algebra is now given by $c_{\mathcal L} = 1 - 6 \hat Q^2$. As we have argued in the previous subsection, it is somewhat natural to introduce the 3-point coupling of this imaginary Liouville theory such that it the shift equation \eqref{3ptse} is satisfied. In terms of the real parameters $\hat \alpha$ and $\hat b$, the shift equation reads \begin{equation} \label{IL3ptse1} \frac{C_{\mathcal{L}}(\hat \alpha_1 - \hat b,\hat \alpha_2,\hat \alpha_3)} {C_{\mathcal{L}}(\hat \alpha_1,\hat \alpha_2,\hat \alpha_3)} \ = \ - \frac{{\gamma}(\hat b(2\hat \alpha_1 - \hat b)){\gamma}(2\hat b\hat \alpha_1)}{\pi \mu_\mathcal{L} {\gamma}(1-\hat b^2) {\gamma}(\hat b(\hat \talpha_{123}-\hat Q))} \frac{{\gamma}(\hat b(\hat \talpha_{23}+\hat b))}{{\gamma}(\hat b\hat \talpha_{13}) {\gamma}(\hat b\hat \talpha_{12})}\ . \end{equation} Here we have simply carried out the substitutions we listed after eqs.\ \eqref{actILiouv} and \eqref{ILcd}. If we shift $\hat \alpha_1$ by $\hat {\beta}$ and invert the relation we obtain, \begin{equation} \label{IL3ptse2} \frac{C_{\mathcal{L}}(\hat \alpha_1 + \hat b,\hat \alpha_2,\hat \alpha_3)} {C_{\mathcal{L}}(\hat \alpha_1,\hat \alpha_2,\hat \alpha_3)} \ = \ - \frac{\pi \mu_\mathcal{L} {\gamma}(\hat b(\hat \talpha_{123}-\hat b^{-1} + 2 \hat b))} {{\gamma}(\hat b^2){\gamma}(\hat b(2\hat \alpha_1)){\gamma}(2\hat b(\hat \alpha_1+\hat b))} \frac{{\gamma}(\hat b(\hat \talpha_{13}+\hat b)) {\gamma}(\hat b(\hat \talpha_{12}+\hat b))} {{\gamma}(\hat b\hat \talpha_{23})}\ . \end{equation} Note that all the factors that depend on linear combination of the variables $\hat \alpha_i$ are the same as in eq.\ \eqref{3ptse}, except for a simple shift by $\hat b$. Factors depending on $\alpha_1$ are not universal since they are effected by the normalization of vertex operators. Following \cite{Zamolodchikov:2005fy} we fix the normalization such that \begin{eqnarray*} \langle \mathcal V_{\hat \alpha}(z_2) \mathcal V_{\hat \alpha}(z_1)\rangle & = & |z_{12}|^{-4 h_{\hat \alpha}} G({\hat \alpha}) \ , \\[2mm] G({\hat \alpha}) &=& \left(\pi \mu_{\mathcal{L}} \gamma(-\hat b^2) \right)^{2{\hat \alpha} \over \hat b} { \gamma(2 {\hat \alpha} \hat b + {\hat b^2}) \, \gamma(2-\hat b^{-2}) \over \gamma(2 + 2{\hat \alpha} \hat b^{-1}- \hat b^{-2}) \, \gamma(\hat b^2) }\ . \end{eqnarray*} The expression on the left hand side is obtained from the second term in eq.\ \eqref{L2pt} by our standard substitutions. Once this normalization is adopted, the associated 3-point couplings take the form \begin{eqnarray} \label{ILstr} C_{\mathcal{L}}({\hat \alpha}_3, {\hat \alpha}_2, {\hat \alpha}_1|\hat b) & = & \left( \pi \mu_\mathcal{L} \gamma\left(-\hat b^2\right) \right)^{{\hat \alpha}_{123} \over \hat b} \, b^{2(b+b^{-1})({\hat \alpha}_{123} +\hat b - \frac{1}{\hat b}) } \frac{\gamma\!\left(2- \hat b^{-2} \right)}{\gamma\!\left(\hat b^{2} \right)}\, \hat b^2\\[2mm] & & \hspace*{-2cm} \times \frac{\Upsilon_{\hat b}\left( {\hat \alpha}_{123} - \hat b^{-1}+2\hat b\right) \Upsilon_{\hat b}\left( {\hat \alpha}_{12} + \hat b \right)\Upsilon_{\hat b}\left( {\hat \alpha}_{23} + \hat b \right)\Upsilon_{\hat b}\left( {\hat \alpha}_{13} + \hat b \right)} {\Upsilon_{\hat b}^0\, \Upsilon_{\hat b}\left( 2 {\hat \alpha}_1 +\hat b \right) \Upsilon_{\hat b}\left( 2 {\hat \alpha}_2 +\hat b \right) \Upsilon_{\hat b}\left( 2 {\hat \alpha}_3 +\hat b \right)} \ . \nonumber \end{eqnarray} It is easy to check that these structure constants solve the shift equations \eqref{IL3ptse2}, though with a different $\alpha_1$-dependent prefactor. This concludes our presentation of imaginary Liouville theory. \section{Bosonization of ${\mathcal N}$=1 Liouville field theory} It is well known \cite{DiFrancesco:1997nk} that a certain orbifold of the product of two real fermions can be bosonized, i.e.\ it is equivalent to a compactified free boson with compactification radius $R =1$. We will now show that a similar bosonization exists for an orbifold of the product of ${\mathcal N}$=1 Liouville field theory with a free fermion $\eta$. In this case, the bosonic description involves two Liouville fields, one with real and the other with imaginary parameter $b$. This relation was first conjectured in \cite{Belavin:2011sw} for the Neveu-Schwarz sector of the supersymmetric Liouville field theory. We will extend the correspondence to the Ramond sector and perform extensive tests for a number of local 3-point functions. \subsection{Product of ${\mathcal N}$=1 Liouville and a fermion} Before we discuss the product of ${\mathcal N}$=1 Liouville theory and a free fermion $\eta$, let us briefly review a few things about a product of fermions. As before, we shall denote one of our fermions by $\psi, \bar\psi$ and the other by $\eta, \bar \eta$. Both $\psi$ and $\eta$ are assumed to possess the same standard operator product, i.e. $$ \psi(z) \psi(w) \sim \frac{1}{z-w} \quad , \quad \eta(z) \eta(w) \sim \frac{1}{z-w} \ . $$ While the first fermion $\psi$ is assumed to be real, i.e.\ $\psi^\dagger = \psi^* = \psi$, we will modify the usual conjugation for $\eta$ such that $\eta^\dagger = -\eta$. In this sense, the fermion $\eta$ may be considered imaginary. Note that the usual conjugation $\eta^* = \eta$ differs from the conjugation $\dagger$ by a simple automorphism of the fermionic theory. In fact, the map $\eta \rightarrow - \eta$ preserves the operator product of the fermion $\eta$. While the algebraic properties of the two fermions are identical, we will use a different bilinear form on their state spaces, one that preserves $\dagger$ rather than the usual $\ast$. As one can easily see, this form is indefinite. Our choice will be motivated a posteriori through the relation with double Liouville theory (see next subsection). Alternatively, one may observe that an imaginary fermion $\eta$ emerges naturally in the reduction from the OSP(1$|$2) WZW model to ${\mathcal N}$=1 Liouville field theory (see formula (2.17) of \cite{Hikida:2007sz}). The theory of a single fermion possesses six conformal primaries, namely the identity, the fermion fields, the energy density and two spin fields. The latter will be denoted by $\varsigma^\pm$ and $\sigma^\pm$ for fermions $\psi$ and $\eta$, respectively. In order to fix our conventions for $\sigma^\pm$, let us state the analogue of the relations \eqref{Rfields} \begin{equation} \eta_0 \sigma^\pm = \frac{1}{\sqrt2} \, e^{\mp i \frac{\pi}{4}} \, \sigma^\mp, \qquad \bar \eta_0 \sigma^\pm = \frac{1}{\sqrt2} \, e^{\pm i \frac{\pi}{4}} \, \sigma^\mp\ . \end{equation} Here, $\eta_0$ and $\bar \eta_0$ denote the zero modes of the fermionic fields $\eta$ and $\bar \eta$, respectively. Due to the conjugation rules of the fermion, i.e.\ $ \eta_{-n}^\dagger = - \eta_{n}$ and $\bar \eta_{-n}^\dagger = - \bar \eta_{n}$, the norms of the R fields satisfy $\langle \sigma^- | \sigma^- \rangle = - \langle \sigma^+ | \sigma^+ \rangle.$ Hence, one of the states $|\sigma^\pm\rangle$ has negative norm. Coming back to the product theory between the fermion $\psi$ and $\eta$, we note that it contains a closed subset of even local fields given by \begin{equation} 1\ , \ \psi\bar \psi\ , \ \psi \eta\ ,\ \psi \bar \eta\ ,\ \bar \psi \eta\ ,\ \bar \psi \bar \eta\ , \ \eta \bar \eta\ ,\ \psi\bar\psi \eta \bar \eta\ ;\ r^\pm = \frac12(\varsigma^+\sigma^+ \pm \varsigma^- \sigma^-)\ . \label{list} \end{equation} The associated conformal blocks give rise to a modular invariant partition function \begin{equation} Z_{\text{\rm fermion}}(q,\bar q) = |\chi_{(0,0)}+ \chi_{(\frac12,\frac12)}|^2 + |\chi_{(0,\frac12)}+ \chi_{(\frac12,0)}|^2 + 2 |\chi_{(\frac{1}{16}, \frac{1}{16})}|^2 \end{equation} where $\chi_{(h,h')} = \chi^1_h \chi^2_{h'}$ are the characters of $c=1/2$ Virasoro representations with lowest weight $h = h_\psi$ and $h' = h'_\eta$. All the fields that are included in $Z_{\text{fermion}}$ can be bosonized through a single bosonic field $Y$ at compactification radius $R = 1$ \cite{DiFrancesco:1997nk}. The exponential fields $\exp(ikY)$ possess a rather simple expression in terms of $\chi = \psi + i \eta$ along with the two spin fields $r^\pm$ we introduced in eq.\ \eqref{list}. For $k \in \mathbb{Z}$ and $k \geq 0$ one finds \begin{eqnarray} :\!e^{ikY}\!:\ = \ :\!\prod_{i=0}^{k-1} \frac{1}{k!}\, \partial^{(i)} \chi \, \bar\partial^{(i)}\bar\chi\!:\ . \label{fermionization1} \end{eqnarray} When $k \in \mathbb{Z}+ \frac12$ we need to use the spin fields $r^\pm$. For $k \geq 1/2$ one has \begin{eqnarray} :\! e^{ikY}\!: \ =\ :\!r^+\!\!\prod_{i=1}^{k-1/2} \frac{1}{k!}\, \partial^{(i)} \chi\, \bar\partial^{(i)}\bar\chi\!: \ . \label{fermionization2} \end{eqnarray} We shall now replace the first fermion by ${\mathcal N}$=1 Liouville field theory with a Liouville field $\varphi$ and a fermion $\psi$. The total central charge of our product theory is \begin{equation}\label{centralN1F} c = c_{SL}+ \frac12 = 2 + 3\left(b + \frac{1}{b}\right)^2 = 8 + 3 b^2 + 3 b^{-2}\ . \end{equation} In our construction of fields we restrict to the even ones, just as for the free fermion model we described above. For $k = 0,1/2,1,3/2, \dots$ we set \begin{equation} \label{Phik} \Phi_\alpha^{(k)}(z,\bar z) = :\exp \left({\alpha \varphi(z,\bar z) + ikY(z,\bar z)}\right):\ . \end{equation} The fields $\Phi_\alpha^{(k)}$ differ from those introduced in \cite{Belavin:2011sw} by their normalization (see also comments below). For negative $k = -1/2,-1,-3/2, \dots$ we introduce $\Phi^{(k)}_\alpha$ through the simple prescription \begin{equation} \Phi_\alpha^{(k)}(z,\bar z) = \Phi_{Q-\alpha}^{(-k)}(z,\bar z) \, =\, :\exp \left({(Q-\alpha) \varphi(z,\bar z) - ikY(z,\bar z)}\right):\ . \end{equation} Up to the normalization we mentioned before, the fields $\Phi_\alpha^{(-|k|)}$ also agree with those defined in \cite{Belavin:2011sw}. The conformal weight of $\Phi_\alpha^{(k)}$ is given by $$ \Delta_\alpha + \frac{k^2}{2} = \frac12 \alpha(Q-\alpha) + \frac{k^2}{2} \ . $$ We note that fields with $|k| \leq 1/2$ are primary with respect to the product of the ${\mathcal N}$=1 super-conformal algebra and the free fermion $\eta$. These primary fields are given by $\Phi^{(0)}_\alpha = \phi_\alpha$ and \begin{equation} \Phi^{(-\frac12)}_\alpha = \left( \sigma^+ \Sigma^+_\alpha - \sigma^- \Sigma^-_\alpha \right) \ , \quad \quad \Phi^{(\frac12)}_\alpha =\frac{1}{2i} \chi_0 \bar\chi_0\Phi^{(-\frac12)}_\alpha = \left(\sigma^+ \Sigma^+_\alpha + \sigma^- \Sigma^-_\alpha \right)\ . \end{equation} For all other values of $k$, the fields $\Phi^{(k)}_\alpha$ are descendent fields. Our explicit computations below will only involve the case of $k = \pm 1, \pm 3/2$. Using the definition \eqref{Phik} and eq.\ \eqref{N1SCA} one can rewrite the first few fields as descendents with respect to the super-conformal algebra and the fermion $\eta$, see also \cite{Belavin:2011sw}, \begin{eqnarray*} \label{Phi1} |\Phi^{(\pm 1)}_\alpha\rangle \!\!\!&=&\! \!\! \Omega^{-2}_{\pm 1}(\alpha) \! \Big[G_{-\frac12} \bar G_{-\frac12}+ (\tfrac{Q}{2} \pm P)^2 \eta_{-\frac12} \bar \eta_{-\frac12} + (\tfrac{Q}{2} \pm P) \!\big( \eta_{-\frac12} \, \bar G_{-\frac12} - \bar \eta_{-\frac12} \, G_{-\frac12} \big)\!\Big]\, |\phi_\alpha\rangle \\[4mm] |\Phi^{(\pm \frac32)}_\alpha\rangle \!\!\! &=& \! \!\! \chi_{-1} \bar\chi_{-1} |\Phi^{(\frac12)}_\alpha\rangle = \Omega_{\pm \frac32}^{-2}(\alpha) \Big[\tfrac{2}{P^{2}} L_{-1}G_0 \bar L_{-1} \bar G_0 + 2 (\tfrac{Q}{2}\pm P)^2 G_{-1} \bar G_{-1} \\[2mm] \!\!\! &+ & \!\!\! \sqrt2 \, \Omega_{\pm \frac32}(\alpha) (\tfrac{Q}{2} \pm P)(\eta_{-1} \bar G_{-1} - \bar \eta_{-1} G_{-1} ) \pm \tfrac{\sqrt2}{P} \Omega_{\pm\frac32}(\alpha) ( \eta_{-1} \bar L_{-1} \bar G_0 - \bar \eta_{-1} L_{-1} G_0) \\[2mm] \!\!\! &+ & \!\!\! \Omega_{\pm\frac32}^2(\alpha) \eta_{-1} \bar \eta_{-1} \pm \tfrac{2}{P}( \tfrac{Q}{2}\pm P)(L_{-1}G_0 \bar G_{-1} + G_{-1} \bar L_{-1} \bar G_0) \Big] \, |\Phi^{(\frac12)}_\alpha\rangle \end{eqnarray*} Here we wrote equations between states rather than fields by means of the usual state-field correspondence. The variable $P$ is related to $\alpha$ through $\alpha = Q/2 + P$. Finally, the pre-factor $\Omega_k(\alpha)$ is given by \begin{equation}\label{Omega} \Omega_k(\alpha) = n_k \prod^{i+j=2|k|}_{\scriptsize \begin{array}{c} i,j = 1 , \\ 2|k| - i - j \in 2 \mathbb{N} \end{array}} (\mbox{\it sign(k)} (2\alpha - Q) + i b + j b^{-1}), \end{equation} where {\it sign}$(k) = k/|k|$ denotes the sign of $k$ when $k \neq 0$ and we set {\it sign}$(0)$=1. The first two constants take the values \begin{equation}\label{nk} n_1 = 2^{-1}, \quad n_{\frac32} = 2^{-\frac32}. \end{equation} There exists a straightforward but cumbersome algorithm that computes the numbers $n_k$ for higher values of $k$. In \cite{Belavin:2011sw}, the factors $\Omega_k$ were absorbed in the normalization of the fields $\Phi^{(k)}_\alpha$. \subsection{Relation with double Liouville theory} We are now prepared to state the main result of this work. It relates the model described in the previous subsection to a product of a Liouville field theory with $c^{(1)} \geq 25$ and an imaginary Liouville theory with $c^{(2)} \leq 1$. We shall often refer to this product as double Liouville theory. According to \cite{Belavin:2011sw}, the $b$-parameters of the two factors must be chosen as \begin{equation} \label{b12} b^{(1)} = { 2b \over \sqrt{2-2b^2} }, \qquad \left( \hat b^{(2)}\right)^{-1} = { 2 \over \sqrt{2-2b^2} }. \end{equation} So that the central charge is $$ c = c_L^{(1)} + c_\mathcal{L}^{(2)} = 2 + 6\left(b^{(1)}+\frac{1}{b^{(1)}}\right)^2 - 6 \left(\hat b^{(2)} - \frac{1}{\hat b^{(2)}}\right)^2 = 8 + 3b^2 + 3 b^{-2}\ .$$ Note that the sum of central charges agrees with the central charge \eqref{centralN1F} of the model we discussed in the previous subsection. Moreover, as was observed in \cite{Crnkovic:1989gy,Crnkovic:1989ug,Lashkevich:1992sb}, the two Virasoro algebras of double Liouville theory can actually be reconstructed from the super-conformal currents $T$ and $G$ along with the fermion $\eta$, \begin{eqnarray} \nonumber L_n^{(1)} &=& \frac{1}{1-b^2} L_n - \frac{1+2b^2}{2- 2b^2} \sum_{r=-\infty}^{\infty} r : \eta_{n-r} \eta_r: + \frac{b}{1-b^2} \sum_{r=-\infty}^{\infty} \eta_{n-r} G_r, \\[-4pt] \label{twoVirasoro} \\[-4pt] \nonumber L_n^{(2)} &=& \frac{1}{1-b^{-2}} L_n - \frac{1+2b^{-2}}{2- 2b^{-2}} \sum_{r=-\infty}^{\infty} r : \eta_{n-r} \eta_r: + \frac{b^{-1}}{1-b^{-2}} \sum_{r=-\infty}^{\infty} \eta_{n-r} G_r. \end{eqnarray} Similar formulas apply to the anti-holomorphic sector, of course. As anticipated in the previous subsection, we now note that the familiar relation $(L^{(i)}_n)^\dagger= L_{-n}^{(i)}$ requires $\dagger$ to act as $\eta^\dagger_n = - \eta_{-n}$ on the modes of the fermion $\eta$. In other words, the Virasoro modes in double Liouville theory possess the usual conjugation rules provided that the modes $L_n$ and $G_n$ do and we take $\eta$ to be imaginary. Given such a close relation between their chiral algebras it seems natural to look for relations between vertex operators. Following \cite{Belavin:2011sw} let us introduce \begin{equation} \label{defVb} \mathbb{V}_\alpha^{(k)}(z,\bar z) = V_{\alpha^{(1)}+kb^{(1)}/2}(z,\bar z) \ \mathcal V_{\hat \alpha^{(2)} + k/2\hat b^{(2)}}(z,\bar z) \end{equation} where $2k$ is an integer, $\alpha$ is a complex parameter and we defined \begin{equation} \alpha^{(1)} = { \alpha \over \sqrt{2-2b^2} } , \qquad \hat \alpha^{(2)} = { b \alpha \over \sqrt{2-2b^2} }. \label{a12} \end{equation} The conformal dimension of the vertex operators \eqref{defVb} is easy to compute with the help of the expressions \eqref{Lcd} and \eqref{ILcd} for conformal weights in (imaginary) Liouville theory, \begin{eqnarray*} h_{(\alpha^{(1)}+kb^{(1)}/2)} + \hat h_{(\hat \alpha^{(2)}+k/2\hat b^{(2)})} = (\alpha^{(1)}+kb^{(1)}/2) (Q^{(1)} - \alpha^{(1)}-kb^{(1)}/2) \\ - (\hat \alpha^{(2)}+ k/2\hat b^{(2)}) (\hat Q^{(2)} - \hat \alpha^{(2)}-k/2\hat b^{(2)}) = \frac12 \alpha (Q-\alpha) + \frac{k^2}{2}\ . \end{eqnarray*} These weights agree with the weights of the fields $\Phi^{(k)}_\alpha$ we introduced in the previous section. Hence, with proper normalizations, the 2-point functions of the fields $\Phi^{(k)}_\alpha$ and $\mathbb{V}^{(k)}_\alpha$ agree. In addition, it is not difficult to check that the fields $\Phi^{(k)}_\alpha$ are primary with respect to Virasoro algebras \eqref{twoVirasoro} of the Liouville field theory and its imaginary cousin. Given these observations it is certainly tempting to contemplate that the relation \begin{equation}\label{main} \Phi^{(k)}_\alpha(z,\bar z) = {\mathcal N}_\alpha^{(k)}\ \mathbb{V}^{(k)}_\alpha(z,\bar z) \end{equation} might hold in arbitrary correlation functions. Through comparison of 3-point functions we shall provide very strong support in favor of this proposal. These computations determine the normalization ${\mathcal N}_\alpha^{(k)}$ to take the form \begin{eqnarray} {\mathcal N}^{( k)}_\alpha &=& (-1)^{k} \, \tilde {\mathcal N}^{( k)}_\alpha, \end{eqnarray} when $k \in \mathbb{N}$, i.e.\ in the NS sector of the theory, and \begin{eqnarray} {\mathcal N}^{( k)}_\alpha &=& 2^{\frac{3}{4}} \, \tilde {\mathcal N}^{( k)}_\alpha, \end{eqnarray} in R sector, i.e.\ when $k$ takes the values $k \in \mathbb{N} + \frac12$. The common factor $\tilde {\mathcal N}^{( k)}_\alpha$ is given by \begin{equation} \tilde {\mathcal N}^{( k)}_\alpha = \frac{\left[\pi \mu_L \gamma\left((b^{(1)})^2\right) \right]^{(\alpha^{(1)} + \frac{k b^{(1)}}{2} ) / b^{(1)}} \left[\pi M \gamma\left(-(\hat b^{(2)})^2 \right) \right]^{-({\hat \alpha}^{(2)} + \frac{k }{2\hat b^{(2)}})/\hat b^{(2)}}} { n_k^2 \, 2^{k^2} \left[\pi \mu \gamma({b Q \over 2}) \right]^{\alpha \over b} b^{- 2k} \left( {1-b^2 \over 2}\right)^{\frac12 + 2k} } . \end{equation} The factors $n_k$ were introduced in eq.\ (\ref{nk}), at least for some special values of $k$. In order to check that the fields $\Phi^{(k)}_\alpha$ and $\mathbb{V}^{(k)}_\alpha$ can be identified in all correlation functions, we must verify that their 3-point functions agree, \begin{equation} \label{3ptprop} \varpi(k_i) \kappa(b) \langle \Phi^{(k_3)}_{\bar \alpha_3} \Phi_{\alpha_2}^{(k_2)} \Phi_{\alpha_1}^{(k_1)} \rangle = {\mathcal N}^{(k_3)}_{\bar \alpha_3} {\mathcal N}^{(k_2)}_{\alpha_2} {\mathcal N}^{(k_1)}_{\alpha_1} \langle \mathbb{V}^{(k_3)}_{\bar \alpha_3} \ \mathbb{V}^{(k_2)}_{\alpha_2} \ \mathbb{V}^{(k_1)}_{\alpha_1}\rangle\ \ , \end{equation} at least up to some constant $\kappa(b)$ that can be absorbed through an appropriate normalization of the vacuum state, see eq.\ \eqref{FUdef} for a concrete formula. The factors $\varpi$ will be shown to satisfy $\varpi^4=1$. Given the complexity of the fields $\Phi^{(k)}$, checking eq.\ \eqref{3ptprop} is a rather non-trivial task. We are not prepared to establish the relation \eqref{3ptprop} for all possible 3-point functions, but we have performed a number of highly non-trivial tests. These are described in the next subsection. \subsection{Comparison of 3-point functions} Our goal is to check relation \eqref{main} in a few selected examples, involving both NS and R sector fields and also super-descendent fields. Most computations are somewhat lengthy but in principle straight forward to carry out. \subsubsection{NS sector} In our first example, we take all three fields of the ${\mathcal N}$=1 Liouville theory to be super-primaries in the NS sector. These are multiplied with the identity field of the free fermion theory, i.e. we consider a 3-point correlator with $\Phi^{(k)}_\alpha =\Phi^{(0)}_\alpha = \phi_\alpha$. Since we have checked already that the conformal dimensions on both sides of the correspondence \eqref{main} match, we shall put the fields at the points $z_3 = \infty$, $z_2 = 1$ and $z_1 = 0$ so that we can omit all dependence on world-sheet coordinates. The 3-point function of $\Phi^{(0)}_\alpha$ is given by $$ \langle \Phi^{(0)}_{\bar \alpha_3} \Phi_{\alpha_2}^{(0)} \Phi_{\alpha_1}^{(0)} \rangle = C_{NS}(\alpha_3,\alpha_2,\alpha_1|b) $$ with $C_{NS}$ as given in equation \eqref{SLstr}. We will use the notation $\bar \alpha_i \equiv Q - \alpha_i $ for reflected momentum of the fields located at infinity. The other side of the correspondence \eqref{main} is given by \begin{eqnarray*} \langle \mathbb{V}^{(0)}_{\bar \alpha_3} \ \mathbb{V}^{(0)}_{\alpha_2} \ \mathbb{V}^{(0)}_{\alpha_1}\rangle &=& C_L(\alpha_3^{(1)}, \alpha_2^{(1)}, \alpha_1^{(1)}| b^{(1)}) \, C_{\mathcal{L}}({\hat \alpha}_3^{(2)}, {\hat \alpha}_2^{(2)}, {\hat \alpha}_1^{(2)}| \hat b^{(2)}). \end{eqnarray*} The arguments $\alpha_i^{(\nu)}$ and $b^{(\nu)}$ that appear in the arguments of the structure constants were introduced in eqs.\ \eqref{a12} and \eqref{b12}. Explicit expressions for the structure constants can be found in eqs.\ \eqref{Lstr} and \eqref{ILstr}. Using the identities \begin{equation} \label{Ups} { \Upsilon_{b^{(1)}}\left( \alpha^{(1)}\right) \over \Upsilon_{\hat b^{(2)}}\left( {\hat \alpha}^{(2)} + \hat b^{(2)}\right)} = B(\alpha) \Upsilon_b^{\rm NS}(\alpha), \end{equation} where \begin{equation} \nonumber \qquad B(\alpha) = {\Upsilon_{b^{(1)}}^0 \over \Upsilon_{\hat b^{(2)}}^0 \Upsilon_{b}^0 } \ b^{b^2 \alpha(Q-\alpha) \over 2 -2b^2} \, \left( 1-b^2 \over 2\right)^{ \alpha(Q-\alpha) -2 \over 4}, \end{equation} stated in (A.9) of \cite{Belavin:2011sw}, one can check that $$ C_L(\alpha_3^{(1)}, \alpha_2^{(1)}, \alpha_1^{(1)}| b^{(1)}) \, C_{\mathcal{L}}({\hat \alpha}_3^{(2)}, {\hat \alpha}_2^{(2)}, {\hat \alpha}_1^{(2)}| \hat b^{(2)}) = A_1 \, C_{NS}(\alpha_3,\alpha_2,\alpha_1|b) $$ with \begin{eqnarray*} A_1 = \frac{2 \left( \pi \mu_L \gamma\!\left({2b^2\over 1-b2}\right) \right)^{Q-\alpha \over 2 b} \!\! \left( \pi M \gamma\!\left({b^2-1 \over 2}\right) \right)^{b a \over 1- b^2}\! \gamma\! \left({- 2 b^2\over 1-b^2} \right) \gamma\! \left({b^2+1 \over 2}\right)}{\left( \left({\pi \mu \over 2}\right) \gamma\!\left({b^2+1 \over 2}\right) \right)^{Q-\alpha \over b} \, \left({2 \over 1-b^2}\right)^{\frac32 }}\, . \end{eqnarray*} Comparison of this $\alpha$-dependent factor with the product of the three normalizations ${\mathcal N}^{(0)}_{\alpha_i}$ we introduced in the previous subsection gives $$ {\mathcal N}^{(0)}_{\bar \alpha_3} {\mathcal N}^{(0)}_{\alpha_2} {\mathcal N}^{(0)}_{\alpha_1} A_1 = \kappa(b) \ . $$ The function $\kappa(b)$ depends on the parameter $b$, but it is independent of the labels $\alpha_i$. Explicitly, it is given by \begin{equation}\label{FUdef} \kappa(b) = \frac{ 2 \left[\pi \mu_L \gamma\left((b^{(1)})^2\right) \right]^{Q^{(1)} \over b^{(1)}} }{ \left[\pi \mu \gamma({b Q \over 2}) \right]^{Q \over b} } \, \gamma\!\left({- 2 b^2\over 1-b^2} \right) \gamma\!\left({b^2+1 \over 2}\right) \ . \end{equation} In conclusion, we have established eq.\ \eqref{3ptprop} for $k_i=0$ with $\varpi(0,0,0) = 1$. Let us now proceed to the next and slightly more complicated example of the relation \eqref{main} in which at least one of the vertex operators involves super-descendents in the ${\mathcal N}$=1 Liouville field. More specifically, let us insert one of the operators $\Phi^{(\pm 1)}_\alpha$ along with two of the operators $\Phi^{(0)}_\alpha$. Looking back at the explicit formulas we spelled out at the end of section 4.1, we observe that only the second term from these expressions can contribute since $\langle \eta \rangle = \langle \bar \eta \rangle = \langle \eta \bar \eta\rangle = 0$. Hence we obtain \begin{eqnarray*} && \hspace{-30pt} \langle \Phi^{(0)}_{\bar \alpha_3} \ \Phi^{( 1)}_{\alpha_2} \ \Phi^{(0)}_{\bar \alpha_1} \rangle = \Omega^{-2}_1(\alpha_2) \langle \phi_{\bar \alpha_3} G_{-\frac12}\bar G_{-\frac12}\phi_{\alpha_2} \phi_{\alpha_1} \rangle = \alpha_2^{-2} \, \tilde C_{NS}(\alpha_3, \alpha_2, \alpha_1|b) \end{eqnarray*} where the evaluation of the correlator in the ${\mathcal N}$=1 Liouville theory uses the structure constants \eqref{SLstr}. On the other side of our correspondence \eqref{main} one finds $$ \langle \mathbb{V}^{(0)}_{\bar \alpha_3} \ \mathbb{V}^{( 1)}_{\alpha_2} \ \mathbb{V}^{(0)}_{\alpha_1} \rangle = C_L(\alpha_3^{(1)}, \alpha_2^{(1)} + \frac{b^{(1)}}{2}, \alpha_1^{(1)}| b^{(1)}) \ C_{\mathcal{L}}({\hat \alpha}_3^{(2)}, {\hat \alpha}_2^{(2)} + \frac{1}{2\hat b^{(2)}}, {\hat \alpha}_1^{(2)}| \hat b^{(2)}) $$ If we insert the explicit formulas \eqref{Lstr} and \eqref{ILstr} for the structure constants $C_L$ and $C_\mathcal{L}$ along with the shift properties \eqref{BGshift} and the identity \begin{equation}\label{Ups2} \Upsilon_R(\alpha) = \frac{b^{b\alpha}}{\gamma\left( \frac{b\alpha+1}{2} \right)} \ \Upsilon_{NS}(\alpha+ b) = B^{-1}(\alpha+b)\, \frac{ b^{b\alpha}} {\left( {1- b^2 \over 2}\right)^{\frac{b\alpha}{2}}} {\Upsilon_{b^{(1)}}\left(\alpha^{(1)} + \frac{b^{(1)}}{2} \right) \over \Upsilon_{\hat b^{(2)}}\left({\hat \alpha}^{(2)} + \frac{1}{2\hat b^{(2)}} + \hat b^{(2)} \right) } \end{equation} where $B(\alpha)$ is the function defined after eq.\ (\ref{Ups}), we can check that \begin{eqnarray*} C_L(\alpha_3^{(1)}, \alpha_2^{(1)}+ \frac{b^{(1)}}{2}, \alpha_1^{(1)}| b^{(1)}) \ C_\mathcal{L}({\hat \alpha}_3^{(2)}, {\hat \alpha}_2^{(2)} + \frac{1}{2\hat b^{(2)}}, {\hat \alpha}_1^{(2)}| \hat b^{(2)}) = A_2 \, \tilde C_{NS}(\alpha_3, \alpha_2, \alpha_1|b) \end{eqnarray*}{equation} where $A_2$ are given by \begin{eqnarray*} A_2 = - \frac{\left( \pi \mu_L \gamma\!\left({2b^2\over 1-b2}\right) \right)^{Q-\alpha-b \over 2 b} \, \left( \pi M \gamma\!\left({b^2-1 \over 2}\right) \right)^{b a +1 \over 1- b^2}\, \gamma\!\left({- 2 b^2\over 1-b^2} \right) \gamma\!\left({b^2+1 \over 2}\right)}{i \left( \left({\pi \mu \over 2}\right) \, \gamma\!\left({b^2+1 \over 2}\right) \right)^{Q-\alpha \over b} b^{2} \, \left( {2 \over 1-b^2}\right)^{\frac72 } \ \alpha_2^{2} } \, . \end{eqnarray*} As in the previous subsection, it is not difficult to see that the functions $A_2$ may be factorized into a product of three $\alpha$-dependent factors ${\mathcal N}$, i.e. $$ {\mathcal N}^{(0)}_{\bar\alpha_3} {\mathcal N}^{( 1)}_{\alpha_2} {\mathcal N}^{(0)}_{\alpha_1} A_2 = i \kappa(b) \, \alpha_2^{-2} \ .$$ The constant $\kappa(b)$ was introduced in eq.\ \eqref{FUdef} above. Combining these results we conclude that eq.\ \eqref{3ptprop} also holds for $k_1 = k_3 = 0$ and $k_2=1$ with the same constant $\kappa(b)$ as in the previous computation and $\varpi(0,1,0)= i$. As a final check for fields from the NS sector we want to consider a correlation function in which two fields have $k=1$, \begin{eqnarray*} && \hspace{-40pt} \langle \Phi^{(0)}_{\bar \alpha_3} \ \Phi^{(1)}_{\alpha_2} \ \Phi^{(1)}_{\alpha_1} \rangle = \Omega_1^{-2}(\alpha_1) \Omega_1^{-2}(\alpha_2) \\[2mm] \!&\times&\! \Big( (\tfrac{Q}{2} + P_1 )^2(\tfrac{Q}{2} + P_2 )^2 \langle \phi_{\alpha_3} \, \eta_{-\frac12} \bar \eta_{-\frac12} \phi_{\alpha_2} \, \eta_{-\frac12} \bar \eta_{-\frac12} \phi_{\alpha_1} \rangle + \langle \phi_{\alpha_3} \, G_{-\frac12}\bar G_{-\frac12}\phi_{\alpha_2} \, G_{-\frac12}\bar G_{-\frac12}\phi_{\alpha_1} \rangle \\[2mm] \!&+&\! ({\tfrac{Q}{2}} + P_1 )(\tfrac{Q}{2} + P_2 )\left( \langle \phi_{\alpha_3} \, \eta_{-\frac12} \bar G_{-\frac12}\phi_{\alpha_2} \, \eta_{-\frac12} \bar G_{-\frac12}\phi_{\alpha_1} \rangle + \langle \phi_3 \, \bar \eta_{-\frac12} G_{-\frac12}\phi_2 \,\bar \eta_{-\frac12} G_{-\frac12}\phi_1 \rangle \right) \Big) \end{eqnarray*} It can be reduced to the basic structure constants with the help of the super-conformal Ward identities \cite{Hadasz:2006qb} \begin{eqnarray*} \langle \varphi_3 \,G_{k}\varphi_2(z,\bar z) \, \varphi_1\rangle &=& \sum\limits_{m=0}^{k+{1\over 2}} \left( \begin{array}{c} \scriptstyle k+{1\over 2}\\[-6pt] \scriptstyle m \end{array} \right) (-z)^{m} \left( \langle G_{m-k}\varphi_3 \,\varphi_2(z,\bar z) \,\varphi_1 \rangle\right. \\[1mm] &&\hspace{20pt} - \epsilon \; \left. \langle \varphi_3 \, \varphi_2(z,\bar z) \,G_{k-m}\varphi_1 \rangle \right), \hskip 5mm k\geqslant-\scriptstyle {1\over 2}, \\[10pt] \langle \varphi_3 \,G_{-k}\varphi_2(z,\bar z) \,\varphi_1\rangle &=& \sum\limits_{m=0}^{\infty} \left( \begin{array}{c} \scriptstyle k-{3\over 2}+m\\[-6pt] \scriptstyle m \end{array} \right) z^{m} \langle G_{k+m} \varphi_3 \,\varphi_2(z,\bar z) \,\varphi_1 \rangle \\[1mm] && \hspace{-95pt} -\; \epsilon (-1)^{k+\frac12 } \sum\limits_{m=0}^{\infty} \left( \begin{array}{c} \scriptstyle k-{3\over 2}+m\\[-6pt] \scriptstyle m \end{array} \right) z^{-k-m+{1\over 2}} \langle \varphi_3 \,\varphi_2(z,\bar z) \, G_{m-{1\over 2}}\varphi_1 \rangle, \hskip 5mm k>\scriptstyle {1\over 2}, \end{eqnarray*} \begin{eqnarray*} \langle G_{-k} \varphi_3 \,\varphi_2(z,\bar z) \,\varphi_1 \rangle = \epsilon \langle \varphi_3\,\varphi_2(z,\bar z) \,G_k\varphi_1 \rangle +\!\!\! \sum\limits_{m=-1}^{l(k-{1\over 2})} \!\!\! \left( \!\! \begin{array}{c} \scriptstyle k+1/2\\[-6pt] \scriptstyle m+1 \end{array} \!\! \right) z^{k-{1\over 2} -m} \langle \varphi_3\,G_{m+{1\over 2}}\varphi_2(z,\bar z) \,\varphi_1 \rangle, \end{eqnarray*} where $\epsilon$ denotes the parity of the field $\varphi_2$ and $l(n) = n$ for $n + 1 \geq 0$ while $l(n) = \infty$ for $n + 1 < 0$. The result reads \begin{eqnarray*} \langle \Phi^{(0)}_{\bar \alpha_3} \ \Phi^{(1)}_{\alpha_2} \ \Phi^{(1)}_{\alpha_1} \rangle \!&=&\! \Omega_1^{-2}(\alpha_1) \Omega_1^{-2}(\alpha_2)\big((\tfrac{Q}{2} + P_1 )^2(\tfrac{Q}{2} + P_2 )^2 + (\Delta_3-\Delta_2 - \Delta_1)^2 \\[2mm] \!&&\! + 2 (\tfrac{Q}{2} + P_1 )(\tfrac{Q}{2} + P_2 ) (\Delta_3-\Delta_2 - \Delta_1) \big) C_{NS}(\alpha_3, \alpha_2, \alpha_1|b) \\[2mm] \!&=&\! \frac{(\tfrac{Q}{2} + P_1 + P_2 -P_3)^2 \, (\tfrac{Q}{2} + P_1 + P_2 +P_3)^2}{4 \alpha_1^2 \, \alpha_2^2 } \, C_{NS}(\alpha_3, \alpha_2, \alpha_1|b) \, . \end{eqnarray*} On the other side we find \begin{eqnarray*} && \hspace{-40pt} \langle \mathbb{V}^{(0)}_{\bar \alpha_3} \ \mathbb{V}^{(1)}_{\alpha_2} \ \mathbb{V}^{(1)}_{\alpha_1} \rangle \\ && = C_L(\alpha_3^{(1)}, \alpha_2^{(1)} + \frac{b^{(1)}}{2}, \alpha_1^{(1)} + \frac{b^{(1)}}{2}| b^{(1)}) \ C_\mathcal{L} ({\hat \alpha}_3^{(2)}, {\hat \alpha}_2^{(2)} + \frac{1}{2\hat b^{(2)}}, {\hat \alpha}_1^{(2)} + \frac{1}{2\hat b^{(2)}}| \hat b^{(2)}) \end{eqnarray*} As before, comparing structure constants and using eqs.\ (\ref{Ups}) and (\ref{Ups2}) we can check that $$ C_L(\alpha_3^{(1)} \! , \alpha_2^{(1)} \! + \! \frac{b^{(1)}}{2} \! , \alpha_1^{(1)} \! + \! \frac{b^{(1)}}{2}| b^{(1)}) C_\mathcal{L}({\hat \alpha}_3^{(2)}\! , {\hat \alpha}_2^{(2)} \! + \! \frac{1}{2\hat b^{(2)}}\! , {\hat \alpha}_1^{(2)} \! + \! \frac{1}{2\hat b^{(2)}}| \hat b^{(2)}) = A_3 C_{NS}(\alpha_3, \alpha_2, \alpha_1|b), $$ where \begin{eqnarray*} A_3 &=& \frac{\left( \pi \mu_L \gamma\!\left({2b^2\over 1-b2}\right) \right)^{Q-\alpha_{123}-2b \over 2 b} \, \left( \pi M \gamma\!\left({b^2-1 \over 2}\right) \right)^{b \alpha_{123} +2 \over 1- b^2}\, \gamma\!\left({- 2 b^2\over 1-b^2} \right) \gamma\!\left({b^2+1 \over 2}\right)}{ \left( \left({\pi \mu \over 2}\right) \, \gamma\!\left({b^2+1 \over 2}\right) \right)^{Q-\alpha_{123} \over b} } \\[4pt] & \times& 2 b^{-4} \, \left( {2 \over 1-b^2}\right)^{-\frac{11}{2} } \, (2\alpha_2)^{-2} (2\alpha_1)^{-2} \, (\alpha_1 + \alpha_2-\alpha_3)^2 \, (\alpha_1 + \alpha_2 + \alpha_3 -Q)^2 . \end{eqnarray*} Thus we have $$ {\mathcal N}^{(0)}_{\bar \alpha_3} {\mathcal N}^{( 1)}_{\alpha_2} {\mathcal N}^{( 1)}_{\alpha_1} A_3 = \kappa(b) \, \frac{(\tfrac{Q}{2} + P_1 + P_2 -P_3)^2 \, (\tfrac{Q}{2} + P_1 + P_2 +P_3)^2}{4 \alpha_2^{2} \, \alpha_1^{2}}. $$ Once more we have established an instance of eq.\ \eqref{3ptprop}, this time for $k_3 = 0$ and $k_1=k_2 = 1$. The constant factor $\kappa(b)$ is given by the same expression as in the previous two cases and $\varpi(0,1,1)=1$. \subsubsection{R sector} So far we have only looked at operators $\Phi^{(k)}_\alpha$ with $k \in \mathbb{Z}$ that involve fields from the NS sector of the ${\mathcal N}$ =1 Liouville field theory. The correspondence \eqref{main} we have formulated also involves fields from the R sector. These appear for values $k \in \mathbb{Z} + \frac12$. Actually, correlators of fields in the R sector have been one of the crucial motivations for this work, see next section. Therefore, we would like to perform a few tests involving $\Phi^{(k)}_\alpha$ with $k \in \mathbb{Z} + \frac12$. The simplest possible 3-point function involving R sector fields is the correlation function \begin{equation} \label{NSRR} \langle \Phi^{(0)}_{\alpha_3} \ \Phi^{(\pm \frac12)}_{\alpha_2} \ \Phi^{(\pm \frac12)}_{\alpha_1} \rangle = \langle \phi_{\alpha_3} \, \sigma^+_2 \Sigma^+_{\alpha_2} \sigma^+_1 \Sigma^+_{\alpha_1} \rangle + \langle \phi_{\alpha_3} \, \sigma^-_2 \Sigma^-_{\alpha_2} \sigma^-_1 \Sigma^-_{\alpha_1} \rangle \end{equation} involving two fields from the R sector along with one from the NS sector. The primary fields $\Sigma^\pm_\alpha$ of the ${\mathcal N}$=1 Liouville field theory are accompanied by the spin fields $\sigma^\pm$ of the free fermion. We added a subscript to these fields in order to keep track of the insertion points. The fields $\sigma^\pm_2$ and $\sigma^\pm_1$ are inserted at $z=1$ and $z=0$, respectively. The field inserted at $z = \infty$ involves the identity field of the free fermion model. Hence, for the 3-point function we consider, we only need to insert 2-point functions of the free fermion model. In passing from the left hand side of eq.\ \eqref{NSRR} we have inserted the definition of $\Phi^{(\pm \frac12)}_\alpha$ and we used that $\langle \sigma^\pm_2 \sigma^\mp_1\rangle = 0$. Assuming that $\sigma^\pm$ have been normalized, the remaining 2-point functions are $\langle \sigma^\pm_2(1) \sigma^\pm_1(0)\rangle = \pm 1$ so that we obtain \begin{equation} \label{NSRR2} \langle \Phi^{(0)}_{\bar \alpha_3} \ \Phi^{(\pm \frac12)}_{\alpha_2} \ \Phi^{(\pm \frac12)}_{\alpha_1} \rangle = \langle \phi_{\bar \alpha_3} \, \Sigma^+_{\alpha_2} \Sigma^+_{\alpha_1} \rangle + \langle \phi_{\bar \alpha_3} \, \Sigma^-_{\alpha_2} \Sigma^-_{\alpha_1} \rangle = 2 \, C^{(+)}_{R}(\alpha_3;\alpha_2,\alpha_1|b) . \end{equation} The relevant correlation functions in ${\mathcal N}$=1 Liouville field theories were spelled out after eq.\ \eqref{SLstrR}. With their help we find \begin{eqnarray*} \nonumber C^{(+)}_{R}(\alpha_3; \alpha_2, \alpha_1|b) \! &=& \! \frac12\left( C^+_{R}(\alpha_3, \alpha_2; \alpha_1|b) + C^-_{R}(\alpha_3; \alpha_2, \alpha_1|b)\right) \\[2mm] \nonumber && \hspace{-40pt} = \frac12 \left( \frac{\pi \mu}{2} \gamma\left(\frac{Qb}{2}\right) b^{1-b^2}\right)^{Q-\talpha_{123} \over b} \ \frac{\Upsilon^0_b \,\Upsilon_b^{\rm R}(2\alpha_1) \Upsilon_b^{\rm R}(2\alpha_2) \Upsilon_b^{\rm NS}(2\alpha_3)}{ \Upsilon_b^{\rm R}(\talpha_{123}-Q) \Upsilon_b^{\rm R}(\talpha_{12}) \Upsilon_b^{\rm NS}(\talpha_{23}) \Upsilon_b^{\rm NS}(\talpha_{13})} \end{eqnarray*} Similarly, we can compute \begin{equation}\label{C(-)} \langle \Phi^{(0)}_{\bar \alpha_3} \ \Phi^{(\frac12)}_{\alpha_2} \ \Phi^{(-\frac12)}_{\alpha_1} \rangle = 2 \, C^{(-)}_R(\alpha_3; \alpha_2, \alpha_1|b) \end{equation} where \begin{eqnarray*} C^{(-)}_{R}(\alpha_3; \alpha_2, \alpha_1|b) \! &=& \! \frac12\left( C^+_{R}(\alpha_3; \alpha_2, \alpha_1|b) - C^-_{R}(\alpha_3; \alpha_2, \alpha_1|b)\right) \\[2mm] \nonumber && \hspace{-40pt} = \frac12 \left( \frac{\pi \mu}{2} \gamma\left(\frac{Qb}{2}\right) b^{1-b^2}\right)^{Q-\talpha_{123} \over b} \frac{\Upsilon^0_b \,\Upsilon_b^{\rm R}(2\alpha_1) \Upsilon_b^{\rm R}(2\alpha_2) \Upsilon_b^{\rm NS}(2\alpha_3)}{ \Upsilon_b^{\rm NS}(\talpha_{123}-Q) \Upsilon_b^{\rm NS}(\talpha_{12}) \Upsilon_b^{\rm R}(\talpha_{23}) \Upsilon_b^{\rm R}(\talpha_{13})} \end{eqnarray*} On the other hand we can compute the 3-point functions of the corresponding fields in double Liouville theory. Using the explicit formulae (\ref{Lstr}) and the relations (\ref{Ups}), (\ref{Ups2}) one may check that \begin{eqnarray*} && \hspace{-20pt} \langle \mathbb{V}^{(0)}_{\bar \alpha_3} \ \mathbb{V}^{(\pm \frac12)}_{\alpha_2} \ \mathbb{V}^{(-\frac12)}_{a_1} \rangle = C_L(\alpha_3^{(1)} \! , \alpha_2^{(1)} \! \pm \! \frac{b^{(1)}}{4} \! , \alpha_1^{(1)} \! - \! \frac{b^{(1)}}{4}| b^{(1)}) C_\mathcal{L} ({\hat \alpha}_3^{(2)} \! , {\hat \alpha}_2^{(2)} \! \pm \! \frac{1}{4\hat b^{(2)}} \! , {\hat \alpha}_1^{(2)} \! - \! \frac{1}{4\hat b^{(2)}}| \hat b^{(2)}) \\ \nonumber && \hspace{80pt} = A_4^\pm C^{(\mp)}_{R}(\alpha_3; \alpha_2,\alpha_1| b) \end{eqnarray*} where \begin{eqnarray*} A_4^\pm \! = \! \frac{2\left( \pi \mu_L \gamma\!\left({2b^2\over 1-b2}\right) \right)^{Q-\alpha_{123} +b/2 \pm b/2 \over 2 b} \! \left( \pi M \gamma\!\left({b^2-1 \over 2}\right) \right)^{b(\alpha_{123}-1/2 \pm 1/2) \over 1- b^2} \!\! \! \gamma\!\left({- 2 b^2\over 1-b^2} \right) \! \gamma\!\left({b^2+1 \over 2}\right) }{\left( \left({\pi \mu \over 2}\right) \, \gamma\!\left({b^2+1 \over 2}\right) \right)^{Q-\alpha \over b} b^{-(1 \mp 1)} \, \left( {2 \over 1-b^2}\right)^{\frac12 \pm 1} } \, . \end{eqnarray*} As in all our previous computations, it is straight forward to show that the functions $A_4^\pm$ may be factorized into a product of three $\alpha$-dependent factors ${\mathcal N}$, up to the familiar $\alpha$-independent term \eqref{FUdef} , i.e. $$ {\mathcal N}^{(0)}_{\bar \alpha_3} {\mathcal N}^{(\pm \frac12)}_{\alpha_2} {\mathcal N}^{(-\frac12)}_{\alpha_1} A^\pm_4 = 2 \kappa(b) \ . $$ Combining these results we conclude once more that eq.\ \eqref{3ptprop} holds, with the familiar $\kappa(b)$ and a factor $\varpi(0,\pm 1/2,-1/2) = 1$. Next let us consider two correlation functions containing the field $\Phi_\alpha^{(\frac{3}{2})}$. In this case in order to express the correlators in terms of the structure constants (\ref{NSRR2}), (\ref{C(-)}) one should use the Ward identities \cite{Friedan:1984rv, Hadasz:2008dt} \begin{eqnarray*} && \hspace{-20pt} \langle G_{-n} \varphi^{\rm R}_3 \, \varphi_2(z, \bar z) \, \varphi_1^{\rm R} \rangle = \epsilon \, \langle \varphi^{\rm R}_3 \, \varphi_2(z, \bar z) \, G_{n} \varphi_1^{\rm R} \rangle + \sum\limits_{k=-\frac12}^{\infty} \Big(\!\! \begin{array}[c]{c} \scriptstyle n+ 1/2 \\[-7pt] \scriptstyle k+1/2 \end{array} \!\!\Big) z^{n-k} \langle \varphi^{\rm R}_3 \, G_{k} \varphi_2(z, \bar z) \,\varphi_1^{\rm R} \rangle, \\[4pt] && \hspace{-20pt} \sum_{p=0}^{\infty} \left(^{\frac12}_p \right) \ z^{\frac12 -p} \ \langle \varphi^{\rm R}_3 \, G_{p-k} \varphi_2(z, \bar z) \,\varphi_1^{\rm R} \rangle = \sum_{p=0}^{\infty} \left(^{ \frac12-k}_{\;\;\;p} \right) (-z)^p \ \langle G_{p+k - \frac12} \varphi^{\rm R}_3 \, \varphi_2(z, \bar z) \,\varphi_1^{\rm R} \rangle \\ &&\hspace{160pt} - \epsilon \sum_{p=0}^{\infty} \left(^{\frac12-k}_{\;\;\;p} \right)(-z)^{ \frac12 -k-p} \langle \varphi^{\rm R}_3 \, \varphi_2(z, \bar z) \, G_{p}\varphi_1^{\rm R} \rangle \, , \end{eqnarray*} where $\varphi_i^{\rm R} $ denotes a R field and $\epsilon$ is the parity of the $NS$ field. Similar Ward identities apply to the fermion $\eta$. With the help of these identities one can see that the simplest correlator with $\Phi^{(\frac32)}$ has only a few non-vanishing terms \begin{eqnarray*} \langle \Phi^{(\pm \frac{1}{2})}_{\bar \alpha_3} \ \Phi^{(0)}_{\alpha_2} \ \Phi^{(\frac32)}_{\alpha_1} \rangle &=& \Omega_{\frac32}^{-2}(\alpha_1) \, \langle \Phi^{(\pm \frac{1}{2})}_{\bar \alpha_3} \Phi^{(0)}_{\alpha_2} \ \Big( 2 P_1^{-2} L_{-1}G_0 \bar L_{-1} \bar G_0 + \frac12 (Q+2P_1)^2 G_{-\frac12} \bar G_{-\frac12} \\[2mm] && + P_1^{-1}(Q+ 2P_1)(L_{-1}G_0 \bar G_{-1}+ G_{-1} \bar L_{-1} \bar G_0 ) \Big) \Phi_{\alpha_1}^{(\frac12)} \rangle, \end{eqnarray*} so that \begin{eqnarray*} \langle \Phi^{(\pm \frac{1}{2})}_{\bar \alpha_3} \Phi^{(0)}_{\alpha_2} \Phi^{(\frac32)}_{\alpha_1} \rangle \!\!\! &=& \!\!\! {i \over \Omega_{\frac32}^{2}(\alpha_1)} \left( (\Delta_3 - \Delta_2 - \Delta_1) - (Q+2P_1)(P_1+P_3)/2 \right)^2 \langle \Phi^{(\pm \frac{1}{2})}_{\bar \alpha_3} \Phi^{(0)}_{\alpha_2} \Phi_{\alpha_1}^{(-\frac12)} \rangle \\[2mm] \!\!\! & = & \!\!\! \frac{ i ( Q + 2P_1 + 2P_2 \pm 2P_3)^2 ( Q + 2P_1 - 2P_2 \pm 2P_3)^2}{ 4 (2P_1 + 2b+ b^{-1})^2 (2P_1 + b+ 2b^{-1})^2} \, C^{(\mp)}(\alpha_2; \alpha_3,\alpha_1| b) \end{eqnarray*} The second correlator is more complicated, \begin{eqnarray*} && \hspace{-20pt} \langle \Phi^{(\pm \frac{1}{2})}_{\bar \alpha_3} \, \Phi^{(1)}_{\alpha_2} \, \Phi^{(\frac32)}_{\alpha_1} \rangle = \Omega_{\frac32}^{-2}(\alpha_1) \Omega_{1}^{-2}(\alpha_2) \\[2mm] && \Big\langle \Phi^{(\pm \frac{1}{2})}_{\bar \alpha_3} \left( G_{-\frac12} \bar G_{-\frac12} + (\tfrac{Q}{2}+P_2)^2 \eta_{-\frac12} \bar \eta_{-\frac12} + (\tfrac{Q}{2} +P_2) (\eta_{-\frac12} \bar G_{-\frac12} - \bar \eta_{-\frac12} G_{-\frac12} \right) \! \Phi^{(0)}_{\alpha_2} \\[2mm] && \ \times \Big( 2^{-1} (Q+2P_1)^2 G_{-1} \bar G_{-1} + 2^{-\frac12} \Omega_{\frac32}(\alpha_1) (Q+2P_1)(\eta_{-1} \bar G_{-1} - \bar \eta_{-1} G_{-1} ) \\[2mm] && \ + 2 {P_1^{-2}} L_{-1}G_0 \bar L_{-1} \bar G_0 + \Omega_{\frac32}^2(\alpha_1) \eta_{-1} \bar \eta_{-1} + \sqrt2 \Omega_{\frac32}(\alpha_1)P_1^{-1} ( \eta_{-1} \bar L_{-1} \bar G_0 - \bar \eta_{-1} L_{-1} G_0) \\[2mm] && \ + P_1^{-1}(Q+2P_1)(L_{-1}G_0 \bar G_{-1} + S_{-1} \bar L_{-1} \bar G_0) \Big) \Phi^{(\frac12)}_{\alpha_1} \Bigg\rangle \ . \end{eqnarray*} Using the Ward identities we arrive at \begin{eqnarray*} && \hspace{-10pt} \langle \Phi^{(\pm \frac{1}{2})}_{\bar \alpha_3} \ \Phi^{(1)}_{\alpha_2} \ \Phi^{(\frac32)}_{\alpha_1} \rangle = \Omega_{\frac32}^{-2}(\alpha_1) \Omega_{1}^{-2}(\alpha_2) \,\Big( (b+2 b^{-1}+2 P_1) \left(2 b+b^{-1}+2 P_1\right) \left(P_2+\tfrac{Q}{2}\right) \\[2mm] &&-2 (P_1\mp P_3) \left(\Delta_3-\Delta_2-\Delta_1-1/2 \right)+2 (2 P_1+Q) (\Delta_3-2\Delta_2- \Delta_1) \\[2mm] &&+2 \left(P_2+\tfrac{Q}{2}\right) (\Delta_3-\Delta_2-\Delta_1)-(P_1\mp P_3) (2 P_1+Q) \left(P_2+\tfrac{Q}{2} \right) \Big)^2 \langle \Phi^{(\pm \frac{1}{2})}_{\alpha_3} \ \Phi^{(0)}_{\alpha_2} \ \Phi^{(\frac12)}_{\alpha_1} \rangle \\[2mm] && = \frac{ (2P_1 + 2P_2 \pm 2P_3+3b + b^{-1})^2 (2P_1 + 2P_2 \pm 2 P_3 +b + 3b^{-1})^2 \ C^{(\pm)}(\alpha_2; \alpha_3,\alpha_1| b)} {8 (2P_1 + 2P_2 \mp 2P_3 +Q)^{-2} (2P_2 + b+ b^{-1})^2 (2P_1 + 2b+ b^{-1})^2 (2P_1 + b+ 2b^{-1})^2 } \end{eqnarray*} Within double Liouville theory we find \begin{eqnarray*} \langle \mathbb{V}^{(\pm \frac12)}_{\bar \alpha_3} \ \mathbb{V}^{(0)}_{\alpha_2} \ \mathbb{V}^{(\frac32)}_{\alpha_1} \rangle &=& C_L(\alpha_3^{(1)}\! \pm \frac{b^{(1)}}{4}, \alpha_2^{(1)} \! , \alpha_1^{(1)} \! + \frac{3b^{(1)}}{4}| b^{(1)}) \, \\[2mm] && C_\mathcal{L}({\hat \alpha}_3^{(2)} \! \pm \frac{1}{4\hat b^{(2)}}, {\hat \alpha}_2^{(2)} \! , {\hat \alpha}_1^{(2)}\! + \frac{3}{4\hat b^{(2)}}| \hat b^{(2)}) = A_5^{\pm} C_{R}^{(\mp)}(\alpha_2; \alpha_3,\alpha_1| b) \\[6pt] \langle \mathbb{V}^{(\pm \frac12)}_{\bar \alpha_3} \ \mathbb{V}^{(1)}_{\alpha_2} \ \mathbb{V}^{(\frac32)}_{\alpha_1} \rangle &=& C_L(\alpha_3^{(1)} \! \pm \frac{b^{(1)}}{4}, \alpha_2^{(1)} \! + \! \frac{b^{(1)}}{2} , \alpha_1^{(1)} \! + \frac{3b^{(1)}}{4}| b^{(1)}) \, \\[2mm] &&\hspace*{-3mm} C_\mathcal{L}({\hat \alpha}_3^{(2)} \! \pm \frac{1}{4\hat b^{(2)}}, {\hat \alpha}_2^{(2)} \! + \frac{1}{2\hat b^{(2)}} , {\hat \alpha}_1^{(2)}\! + \frac{3}{4\hat b^{(2)}}| \hat b^{(2)}) = A_6^{\pm} C_{R}^{(\pm)}(\alpha_2; \alpha_3,\alpha_1| b) \end{eqnarray*} where \begin{eqnarray*} A_5^{\pm} \!\! & = & \!\! \frac{ \left( \pi \mu_L \gamma\!\left({2b^2\over 1-b2}\right) \right)^{Q-\alpha - 3b/2 \mp b/2 \over 2 b} \, \left( \pi M \gamma\!\left({b^2-1 \over 2}\right) \right)^{b(a+ 3/2 \pm 1/2)\over 1- b^2} \gamma\!\left({- 2 b^2\over 1-b^2} \right) \gamma\!\left({b^2+1 \over 2}\right) }{ \left( \left({\pi \mu \over 2}\right) \, \gamma\!\left({b^2+1 \over 2}\right) \right)^{Q-\alpha \over b} \, b^{3\pm 1} \, \left( {2 \over 1-b^2}\right)^{\frac{9}{2}\pm 1 } } \, \\[6pt] && \hspace{-20pt} \frac{ ( Q + 2P_1 + 2P_2 \pm 2P_3)^2 ( Q + 2P_1 - 2P_2 \pm 2P_3)^2}{ 8 \, (2P_1 + 2b+ b^{-1})^2 (2P_1 + b+ 2b^{-1})^2} \\[8pt] A_6^{\pm} \!\! & = & \!\! -\frac{ \left( \pi \mu_L \gamma\!\left({2b^2\over 1-b2}\right) \right)^{Q-\alpha - 5b/2 \mp b/2 \over 2 b} \, \left( \pi M \gamma\!\left({b^2-1 \over 2}\right) \right)^{b(a+ 5/2 \pm 1/2)\over 1- b^2} \gamma\!\left({- 2 b^2\over 1-b^2} \right) \gamma\!\left({b^2+1 \over 2}\right) }{ \left( \left({\pi \mu \over 2}\right) \, \gamma\!\left({b^2+1 \over 2}\right) \right)^{Q-\alpha \over b} b^{5 \pm 1} \, \left( {2 \over 1-b^2}\right)^{\frac{13}{2}\pm 1 } } \\[6pt] && \hspace{-20pt} \frac{(2P_1 + 2P_2 \mp 2P_3 +Q)^2 (2P_1 + 2P_2 \pm 2P_3 +3b + b^{-1})^2 (2P_1 + 2P_2 \pm 2 P_3 +b + 3b^{-1})^2 }{32 \, (2P_2 + b+ b^{-1})^2 (2P_1 + 2b+ b^{-1})^2 (2P_1 + b+ 2b^{-1})^2 } \end{eqnarray*} Comparing with the correlators from the first part of the computation in ${\mathcal N}$=1 Liouville theory we obtain, \begin{eqnarray*} && \hspace{-10pt} {\mathcal N}^{(\pm \frac12)}_{\bar \alpha_3} {\mathcal N}^{(0)}_{\alpha_2} {\mathcal N}^{(\frac32)}_{\alpha_1} A^\pm_5 = 2 \kappa(b) \, \frac{ ( Q + 2P_1 + 2P_2 \pm 2P_3)^2 ( Q + 2P_1 - 2P_2 \pm 2P_3)^2}{ 8 \, (2P_1 + 2b+ b^{-1})^2 (2P_1 + b+ 2b^{-1})^2}\ , \\[4pt] && \hspace{-10pt}{\mathcal N}^{(\pm \frac12)}_{\bar \alpha_3} {\mathcal N}^{(1)}_{\alpha_2} {\mathcal N}^{(\frac32)}_{\alpha_1} A^\pm_6 = 4 \kappa(b) \\[2mm] && \frac{(2P_1 + 2P_2 \mp 2P_3 +Q)^2 (2P_1 + 2P_2 \pm 2P_3 +3b + b^{-1})^2 (2P_1 + 2P_2 \pm 2 P_3 +b + 3b^{-1})^2 }{32 \, (2P_2 + b+ b^{-1})^2 (2P_1 + 2b+ b^{-1})^2 (2P_1 + b+ 2b^{-1})^2 } \end{eqnarray*} so that we verified two additional cases of eq.\ \eqref{3ptprop} with $\varpi(\pm 1/2,0,3/2)= -i$ and $\varpi(\pm 1/2,1,3/2) = 1$. This concludes the tests of our main correspondence \eqref{main}. \section{Outlook and Conclusions} The main result of this work is our formula \eqref{main} that relates fields in the product of ${\mathcal N}$=1 Liouville field theory with a free fermion to primaries in double Liouville field theory. We have tested this proposal through a number of non-trivial calculations. The correspondence \eqref{main} extends related observations in \cite{Belavin:2011sw} to the R sector. In addition, we have been able to normalize the fields in both R and NS sector such that the 3-point functions agree up to a simple $b$-dependent factor $\sim \kappa$. Since this factor does not depend on the fields we insert, it can be absorbed through a normalization of the vacuum state. Our results may be extended in a number of different directions. It clearly seems worthwhile to study the correspondence \eqref{main} for correlation functions e.g.\ on discs with non-trivial boundary conditions or higher genus surfaces. ${\mathcal N}$ =1 Liouville field theory possesses one continuous family of boundary conditions which preserve the ${\mathcal N}$ =1 super-conformal algebra. Though boundary conditions in imaginary Liouville theory have not received as much attention as the bulk model, see however \cite{Fredenhagen:2004cj}, it seems likely that double Liouville theory admits conformal boundary conditions that are parametrized by two continuous labels. A subset of these boundary conditions should preserve the larger ${\mathcal N}$=1 super-conformal symmetry along with simple gluing conditions for the fermion $\eta$. In an interesting recent paper \cite{Gaiotto:2012wh} Gaiotto engineers a conformal interface between the minimal models MM$_k$ and MM$_{k-1}$. Gaiotto's construction makes essential use of the relation \eqref{mainrat} between the product theory and supersymmetric minimal models. It seems likely that a similar interface between Liouville theory and its imaginary version also exists. Constructing this interface explicitly might be of some interest as it could provide more insight into the relation between standard Liouville field theory and its imaginary cousin. The main motivation for this work, however, came from the results of \cite{Hikida:2007sz} which relate correlators of the OSP(1$|$2) WZW model at level $k$ to those of ${\mathcal N}$=1 Liuoville theory with $b^{-2} = 2k-3$. In order to compute N-point functions of primaries $V^\epsilon_j(\mu|z)$ in the WZW model, one needs to calculate higher correlators in a product of ${\mathcal N}$=1 Liouville field theory with a free fermion. The latter involve N fields from the physical spectrum of the supersymmetric Liouville theory along with N-2 degenerate ones whose insertion points $y_i = y_i (\mu_\nu)$ depend on the complex parameters $\mu_\nu$. It turns out that all these fields must be takes from the R sector of the model. More precisely, one finds \begin{equation}\label{OSP} \langle \prod_{\nu=1}^N V^{\epsilon_\nu}_{j_\nu}(\mu_\nu | z_\nu)\rangle \sim \delta^2(\sum_{\nu=1}^N \mu_\nu) \langle \prod_{\nu=1}^N \frac12 \left( \Phi^{(-\frac12)}_{\alpha_\nu} (z_\nu) - i \epsilon_\nu \Phi^{(\frac12)}_{\alpha_\nu} (z_\nu) \right) \prod_{j=1}^{N-2} \Phi^{(\frac12)}_{- \frac{1}{2b}} (y_j) \rangle\ . \end{equation} up to some simple factors. In the present article we have argued that the correlation functions on the right hand side can be calculated in double Liouville theory. We believe that such a relation between the OSP(1$|$2) WZW model and double Liouville theory could become a crucial ingredient in finding a supersymmetric analogue of the celebrated FZZ-duality between the SL(2)$/$U(1) black hole sigma model and sine-Liouville field theory, much along the lines of \cite{Hikida:2008pe}. In this context it is crucial to observe that, according to eq.\ \eqref{main}, the degenerate fields we have to insert at points $y_j$ on the right hand side of the correspondence \eqref{OSP} are trivial in imaginary Liouville theory, i.e.\ $$ \Phi^{(\frac12)}_{- \frac{1}{2b}} (y) \sim V_{-\frac{1}{2b}}(y)\ . $$ Hence, the imaginary Liouville theory is merely a spectator throughout most of the computations performed in \cite{Hikida:2008pe}. Consequently, we can express correlation functions in the OSP(1$|$2) WZW model through a product of sine-Liouville and imaginary Liouville theory. It then remains to rewrite the latter in terms of a more conventional theory. We shall return to these issues in a forthcoming paper. \bigskip \noindent {\bf Acknowledgements:} We wish to thank Stefan Fredenhagen, Leszek Hadasz, Yasuaki Hikida, Zbigniew Jaskolski and J\"org Teschner for useful discussions and comments. This work was supported in part by the SFB 676. The work of PS was supported by the Kolumb Programme KOL/6/2011-I of FNP and by the NCN grant DEC2011/01/B/ST1/01302.
\section[#1]{#2}} \def\pr {\noindent {\it Proof.} } \def\rmk {\noindent {\it Remark} } \def\n{\nabla} \def\bn{\overline\nabla} \def\ir#1{\mathbb R^{#1}} \def\hh#1{\Bbb H^{#1}} \def\ch#1{\Bbb {CH}^{#1}} \def\cc#1{\Bbb C^{#1}} \def\f#1#2{\frac{#1}{#2}} \def\qq#1{\Bbb Q^{#1}} \def\cp#1{\Bbb {CP}^{#1}} \def\qp#1{\Bbb {QP}^{#1}} \def\grs#1#2{\bold G_{#1,#2}} \def\bb#1{\Bbb B^{#1}} \def\dd#1#2{\frac {d\,#1}{d\,#2}} \def\dt#1{\frac {d\,#1}{d\,t}} \def\mc#1{\mathcal{#1}} \def\nA#1{\big|\nabla|A^{#1}|^2\big|} \def\pr{\frac {\partial}{\partial r}} \def\pfi{\frac {\partial}{\partial \phi}} \def\pf#1{\frac{\partial}{\partial #1}} \def\pd#1#2{\frac {\partial #1}{\partial #2}} \def\ppd#1#2{\frac {\partial^2 #1}{\partial #2^2}} \def\epw#1{\ep_1\w\cdots\w \ep_{#1}} \def\td{\tilde} \def\ul{\underline} \font\subjefont=cmti8 \font\nfont=cmr8 \def\inner#1#2#3#4{(e_{#1},\ep_1)(e_{#2},\ep_2)(\nu_{#3},\ep_1)(\nu_{#4},\ep_2)} \def\second#1#2{h_{\a,i#1}h_{\be,i#2}\lan e_{#1\a},A\ran\lan e_{#2\be},A\ran} \def\a{\alpha} \def\be{\beta} \def\gr{\bold G_{2,2}^2} \def\r{\Re_{I\!V}} \def\sc{\bold C_m^{n+m}} \def\sg{\bold G_{n,m}^m(\bold C)} \def\p#1{\partial #1} \def\pb#1{\bar\partial #1} \def\de{\delta} \def\De{\Delta} \def\e{\eta} \def\ep{\varepsilon} \def\eps{\epsilon} \def\G{\Gamma} \def\g{\gamma} \def\k{\kappa} \def\la{\lambda} \def\La{\Lambda} \def\om{\omega} \def\Om{\Omega} \def\th{\theta} \def\Th{\Theta} \def\si{\sigma} \def\Si{\Sigma} \def\ul{\underline} \def\w{\wedge} \def\vs{\varsigma} \def\Hess{\mbox{Hess}} \def\R{\Bbb{R}} \def\C{\Bbb{C}} \def\tr{\mbox{tr}} \def\U{\Bbb{U}} \def\lan{\langle} \def\ran{\rangle} \def\ra{\rightarrow} \def\Dirac{D\hskip -2.9mm \slash\ } \def\dirac{\partial\hskip -2.6mm \slash\ } \def\bn{\bar{\nabla}} \def\aint#1{-\hskip -4.5mm\int_{#1}} \def\V{\mbox{Vol}} \def\ol{\overline} \def\ze{\zeta} \renewcommand{\subjclassname} \textup{2000} Mathematics Subject Classification} \subjclass{58E20,53A10.} \begin{document} \title [Curvature estimates] {Curvature estimates for minimal submanifolds of higher codimension and small G-rank} \author [J. Jost, Y. L. Xin and Ling Yang]{J. Jost, Y. L. Xin and Ling Yang} \address{Max Planck Institute for Mathematics in the Sciences, Inselstr. 22, 04103 Leipzig, Germany, and \\ Department of Mathematics, University of Leipzig, 04081 Leipzig, Germany.} \email{<EMAIL>} \address {Institute of Mathematics, Fudan University, Shanghai 200433, China.} \email{<EMAIL>} \address{Institute of Mathematics, Fudan University, Shanghai 200433, China.} \email{<EMAIL>} \thanks{The first author is supported by the ERC Advanced Grant FP7-267087. The second author and the third author are supported partially by NSFC. They are also grateful to the Max Planck Institute for Mathematics in the Sciences in Leipzig for its hospitality and continuous support. } \begin{abstract} We obtain new curvature estimates and Bernstein type results for minimal $n-$submanifolds in $\ir{n+m},\, m\ge 2$ under the condition that the rank of its Gauss map is at most 2. In particular, this applies to minimal surfaces in Euclidean spaces of arbitrary codimension. \end{abstract} \maketitle \Section{Introduction}{Introduction} The classical Bernstein theorem says that an entire minimal graph in $\ir{3}$ has to be an affine plane. The mathematics behind this theorem has proved to be enormously rich. It connects with differential geometry, partial differential equations, and complex analysis, and it has been a stimulus for important developments in all these fields (for some references, see for instance \cite{x}). In particular, the question emerged and has been intensively investigated to what extent this result can be generalized in various directions, that is, under which conditions a minimal submanifold of some Euclidean space (or a sphere) is necessarily affine linear (or a sub-sphere). In particular, Bernstein type theorems for higher dimension and codimension have been studied. Thus, let $M\to\ir{n+m}$ be an $n$ dimensional submanifold in Euclidean space $\ir{n+m}$. In recent work (\cite{j-x}, \cite{j-x-y2} and \cite{j-x-y3}), we have systematically used geometric properties of Grassmannian manifolds and the regularity theory of harmonic maps to obtain new Bernstein type results for higher dimension $n\ge 3$ and codimension $m\ge 2$. The key point is that the Gauss map of such a minimal submanifold is a harmonic map with values in a Grassmann manifold. Thus, our approach combines methods from differential geometry and partial differential equations. This leads us to the question whether this can also be combined with the complex analysis approach. The complex analysis approach is, of course, naturally restricted to the case $n=2$, if, for the sake of the discussion, we ignore such issues as subvarieties of complex spaces. Thus, the present paper is concerned with the case $n=2$ and $m\ge 2$. Now, the target manifold of the Gauss map is $\grs{2}{m}$, the complex quadric, and the Gauss map is holomorphic. A powerful traditional approach to this problem investigates the value distribution of the Gauss image within the framework of complex geometry. This was started by Chern and Osserman \cite{c-o}. From their results, an analogue of Moser's Bernstein theorem \cite{m} that works for $n\ge 2$ and $m=1$ also holds for the case $n=2$ and $m\ge 2$. More precisely, for an entire minimal graph given by $f:\ir{2}\to\ir{m}$ if $\De_f=[\det(\de_{ij}+f^\a_if^\a_j)]^{\f{1}{2}}$ is uniformly bounded, then $f$ is affine linear, and thus represents an affine plane in $\ir{2+m}$. In the present paper, we use curvature estimate techniques to improve this result. This will also enable us to achieve some technical generalization which we shall now formulate. For a minimal $n-$submanifold $M$ in $\ir{n+m}$ we consider the rank of the Gauss map, which is called the $G-rank$ for simplicity. Our condition then simply is $G-rank\le 2$. Obviously, this class of submanifolds contains surfaces in $\ir{2+m}$, as well as cylinders over surfaces in $\ir{3}$. In \cite{d-f}, Dajczer and Florit gave a parametric description of all Euclidean minimal submanifolds of $G-rank = 2$. In particular, they showed that complete minimal submanifolds with $G-rank=2$ have dimension $n=3$ at most (without Euclidean factor). Here, then, are our main results. \begin{thm} Let $M$ be an $n$-dimensional complete minimal submanifold in $\R^{n+m}$ with $G-rank\le 2$ and positive $w$-function (see (\ref{wf}). If $M$ has polynomial volume growth and the function $v=w^{-1}$ has growth \begin{equation} \max_{D_R(p_0)}v=o(R^{\f{2}{3}}) \end{equation} for a fixed point $p_0$, then $M$ has to be an affine linear subspace. \end{thm} Then, we have \begin{thm}\label{t3} Let $M=\{(x,f(x)):x\in \R^n\}$ be an entire minimal graph given by a vector-valued function $f:\R^n\ra \R^m$ with $G-rank\le 2$. If the slope of $f$ satisfies \begin{equation} \De_f=\left[\det\Big(\de_{ij}+\sum_{\a}\f{\p f^\a}{\p x^i}\f{\p f^\a}{\p x^j}\Big)\right]^{\f{1}{2}}=o(R^{\f{2}{3}}), \end{equation} where $R^2=|x|^2+|f(x)|^2$, then $f$ has to be an affine linear function. \end{thm} \begin{thm}\label{t4} Let $f:\R^2\ra \R^m$ $(x^1,x^2)\mapsto (f^1,\cdots,f^m)$ be an entire solution of the minimal surface system \begin{equation}\label{mi} \Big(1+\Big|\f{\p f}{\p x^2}\Big|^2\Big)\f{\p^2 f}{(\p x^1)^2}-2\Big\lan \pd{f}{x^1},\pd{f}{x^2}\Big\ran\f{\p^2 f}{\p x^1\p x^2}+\Big(1+\Big|\pd{f}{x^1}\Big|^2\Big)\f{\p^2 f}{(\p x^2)^2}=0. \end{equation} If there exists $\ep>0$, \begin{equation} \De_f=\det\Big(\de_{ij}+\sum_\a \pd{f^\a}{x^i}\pd{f^\a}{x^j}\Big)^{\f{1}{2}}=O(R^{1-\ep}) \end{equation} with $R=|x|$, then $f$ has to be affine linear. \end{thm} This is an improvement of the Chern-Osserman theorem mentioned above which required $\De_f$ to be bounded. The paper is organized as follows. After \S 2 on basic notation and formulae we describe the special features of the $G-rank\le 2$ case in \S 3. Then in \S 4 , using subharmonic functions obtained from the geometry of the Grassmann manifolds and a Bochner type formula for the squared norm of the second fundamental form $|B|^2$ we can obtain $L^p-$estimates and point-wise estimates for $|B|^2$. Those estimates lead to Bernstein type results. \S 5 is devoted to the graphic situation. In the final section we discuss the sharpness of our estimates. We find that holomorphic curves reach all the possible equalities in all the geometric inequalities (\ref{b4}), (\ref{dew}) and (\ref{kato}). We thank Marcos Dajczer for informing us about \cite{d-f}. \Section{Fundamental formulas}{Fundamental formulas}\label{s1} Let $M$ be an $n$-dimensional submanifold in an $(n+m)$-dimensional Riemannian manifold $\bar{M}$ with second fundamental form $B.$ Let $\bar{\n}$ denote the Levi-Civita connection on $\bar{M}$. It naturally induces connections on the tangent bundle $TM$, the normal bundle $NM$ and various induced bundles over $M.$ For notational simplicity all of them are denoted by $\n$. For arbitrary $\nu\in \G(NM)$ the shape operator $A^\nu: TM\ra TM$ satisfies $\lan B_{X,Y},\nu\ran=\lan A^\nu(X),Y\ran$ for every $X,Y\in \G(TM)$. It is self adjoint in the tangent spaces of $M.$ The mean curvature field $H$ is defined to be the trace of the second fundamental form. $M$ is called a \textit{minimal submanifold} whenever $H$ vanishes on $M$ everywhere. The second fundamental form, the curvature tensor of the submanifold, the curvature tensor of the normal bundle and that of the ambient manifold are connected by the Gauss equations, the Codazzi equations and the Ricci equations (see \cite{x}, \S 1.1). In this paper we consider a minimal submanifold $M$ of dimension $n$ in the Euclidean space $\R^{n+m}$ with codimension $m\geq 2$. Now, there is an important tool, the Gauss map. The Gauss map $\g: M\ra \grs{n}{m}$ is defined by $$\g(p)=T_p M\in \grs{n}{m}$$ via the parallel translation in $\R^{n+m}$ for every $p\in M$, where $\grs{n}{m}$ is the Grassmann manifold consisting of the oriented linear $n$-subspaces in $\R^{n+m}$. One can write $$\g(p)=e_1\w \cdots\w e_n$$ by using Pl\"{u}cker coordinates. Here and in the sequel, $\{e_i\}$ is a local orthonormal tangent frame field of $M$ and $\{\nu_\a\}$ denotes a local orthonormal normal frame field of $M$; we use the summation convention and agree on the following ranges of indices: $$1\leq i,j,k\leq n;\qquad 1\leq \a,\be,\g\leq m.$$ Let $$h_{\a,ij}:=\lan B_{e_i e_j},\nu_\a\ran$$ be the coefficients of the second fundamental form $B$ of $M$ in $\R^{n+m}.$ Then, \begin{equation}\label{Gm1} \g_* e_i=h_{\a,ij}e_{j\a}, \end{equation} where $e_{j\a}$ is obtained by replacing $e_j$ by $\nu_\a$ in $e_1\w\cdots\w e_n$. The energy density of the Gauss map thus is nothing but the squared norm of the second fundamental form (see \cite{x} \S 3.1), $$e(\g)=\f{1}{2}\lan \g_*e_i,\g_* e_i\ran=\f{1}{2}\sum_{\a,i,j}h_{\a,ij}^2=\f{1}{2}|B|^2.$$ Given two unit $n$-vectors $$A=a_1\w \cdots\w a_n,\qquad B=b_1\w \cdots\w b_n$$ in the Grassmann manifold $\grs{n}{m}$, their inner product is defined by \begin{equation} \lan A,B\ran=\det\big(\lan a_i,b_j\ran\big) \end{equation} Fixing a simple unit $n$-vector $A=\ep_1\w \cdots\w \ep_n$, we define the $w$-function on $M$: \begin{equation}\label{wf} w(p):=\lan e_1\w\cdots\w e_n,A\ran=\det\big(\lan e_i,\ep_j\ran\ran\big). \end{equation} Via the Pl\"{u}cker imbedding, the Grassmann manifold $\grs{n}{m}$ can be viewed as a submanifold in a Euclidean space, and the $w$-function can be regarded as the composition of the Gauss map and a height function on $\grs{n}{m}$ (see \cite{x-y1}, \cite{j-x-y2} and \cite{j-x-y3}). In particular, if $M=\big(x,f(x)\big)$ is a graph in $\R^{n+m}$ given by a vector-valued function $f:\R^n\ra \R^m$, then choosing $A$ to be one representing $(x_1,\cdots,x_n)$ coordinate $n$-plane implies $w>0$ and moreover \begin{equation} v:=w^{-1} \end{equation} equals the volume element of $M$ (see \cite{j-x}). The Codazzi equations yield the following formulas for the $w$-function: \begin{lem}\cite{fc}\cite{x1} If $M$ is a submanifold in $\R^{n+m}$, then \begin{equation}\label{dw} \n_{e_i} w=h_{\a,ij}\lan e_{j\a},\ep_1\w\cdots\w \ep_n\ran. \end{equation} Moreover if $M$ has parallel mean curvature, i.e. $\n H\equiv 0$, then \begin{equation}\label{La} \De w=-|B|^2 w+\sum_i\sum_{\a\neq \be,j\neq k}h_{\a,ij} h_{\be,ik} \lan e_{j\a,k\be},\ep_1\w \cdots\w \ep_n\ran. \end{equation} with \begin{equation} e_{j\a,k\be}=e_1\w\cdots\w \nu_\a \w\cdots\w \nu_\be\w \cdots\w e_n \end{equation} that is obtained by replacing $e_j$ by $\nu_\a$ and $e_k$ by $\nu_\be$ in $e_1\w \cdots\w e_n$, respectively. \end{lem} To have the curvature estimates we need the Simons' version of the Bochner type formula for the squared norm of the second fundamental form. A straightforward calculation shows (see \cite{Si}, (2.6) in \cite{x1}) \begin{equation}\label{b4} \De |B|^2=2|\n B|^2+2\lan \n^2 B,B\ran\geq 2|\n B|^2-3|B|^4. \end{equation} \Section{Small G-rank cases}{Small G-rank cases} The rank of the Gauss map for a submanifold $M$ in $\ir{n+m}$ is closely related to rigidity problems. The classical Beez-Killing theorem is a local rigidity property for hypersurfaces in $\ir{n+1}$ when $G-rank\ge 3$. For global investigations, we refer to \cite{d-g}. Here, we study the case of $G-rank\le 2$. Now, for every $p\in M$, we have $$\dim \text{Ker}(\g_*)_p=n-\text{rank}(\g_*)_p\geq n-2.$$ Then for any $p_0\in M$, there exists a local smooth distribution $\mc{K}$ of dimension $n-2$ on $U\ni p_0$, such that $\mc{K}_p\subset \text{Ker}(\g_*)_p$ for any $p\in U$. $\mc{K}$ is called the relative nullity distribution by Chern-Kuiper \cite{c-k}. This is an integrable distribution. Therefore, one can find a local tangent orthonormal frame field $\{e_i\}$, such that $\mc{K}_p=\text{span}\{e_i(p):i\geq 3\}$, i.e. \begin{equation} \g_*e_i=0\qquad 3\leq i\leq n \end{equation} and it follows from (\ref{Gm1}) that $h_{\a,ij}=0$, i.e. \begin{equation} B_{e_i e_j}=0\qquad \text{whenever }i\geq 3\text{ or }j\geq 3. \end{equation} Hence \begin{equation} 0=H=\sum_{i=1}^n B_{e_i e_i}=B_{e_1 e_1}+B_{e_2 e_2}. \end{equation} At the considered point, let \begin{equation} G(e_1,e_2):=\left(\begin{array}{cc} \lan B_{e_1 e_1},B_{e_1 e_1}\ran & \lan B_{e_1 e_1},B_{e_1 e_2}\ran\\ \lan B_{e_1 e_2},B_{e_1,e_1}\ran & \lan B_{e_1 e_2},B_{e_1 e_2}\ran \end{array}\right). \end{equation} $G$ then is a semi-positive definite matrix, whose eigenvalues are denoted by $\mu_1^2$ and $\mu_2^2$ ($\mu_1\geq \mu_2\geq 0$) . Then there exists an orthogonal matrix \begin{equation}\label{o1} O=\left(\begin{array}{cc} \cos\th & -\sin\th\\ \sin\th & \cos\th \end{array}\right) \end{equation} such that \begin{equation}\label{o2} G=O\left(\begin{array}{cc} \mu_1^2 & \\ & \mu_2^2 \end{array}\right)O^T. \end{equation} Now we put \begin{equation} f_1=\cos\a e_1-\sin\a e_2\qquad f_2=\sin\a e_1+\cos\a e_2 \end{equation} with $\a$ to be chosen, then $$\aligned B_{f_1f_1}&=\cos^2\a B_{e_1e_1}+\sin^2\a B_{e_2e_2}-2\cos\a\sin\a B_{e_1e_2}\\ &=\cos(2\a) B_{e_1e_1}-\sin(2\a) B_{e_1e_2}\\ B_{f_1f_2}&=\cos\a\sin\a B_{e_1e_1}-\cos\a\sin\a B_{e_2e_2}+(\cos^2\a-\sin^2\a)B_{e_1e_2}\\ &=\sin(2\a) B_{e_1e_1}+\cos(2\a) B_{e_1e_2}. \endaligned$$ Thus \begin{equation}\label{o3} G(f_1,f_2)=\left(\begin{array}{cc} \cos(2\a) & -\sin(2\a)\\ \sin(2\a) & \cos(2\a) \end{array}\right)G(e_1,e_2)\left(\begin{array}{cc} \cos(2\a) & -\sin(2\a)\\ \sin(2\a) & \cos(2\a) \end{array}\right)^T \end{equation} Choosing $\a=-\f{\th}{2}$ and combining with (\ref{o1}), (\ref{o2}) and (\ref{o3}) gives $$G(f_1,f_2)=\left(\begin{array}{cc} \mu_1^2 & \\ & \mu_2^2 \end{array}\right). $$ Therefore, by carefully choosing local tangent frames and normal frames, one can assume that at the considered point \begin{equation}\label{sf} A^1=\left(\begin{array}{cccc} \mu_1 & 0 & &\\ 0 & -\mu_1 & &\\ & & &\\ & & \multicolumn{1}{c}{\raisebox{0.5ex}[0pt]{\Huge o}}& \end{array} \right)\qquad A^2=\left(\begin{array}{cccc} 0 & \mu_2 & &\\ \mu_2 & 0 & &\\ & & &\\ & & \multicolumn{1}{c}{\raisebox{0.5ex}[0pt]{\Huge o}}& \end{array}\right) \end{equation} and $A^\a=0$ for each $\a\geq 3$, where $A^\a:=A^{\nu_\a}$ is the shape operator. In this case, (\ref{La}) can be rewritten as \begin{equation}\label{La2}\aligned \De w&=-|B|^2 w+2\sum_i\sum_{j\neq k}h_{1,ij}h_{2,ik}\lan e_{j1,k2},A\ran\\ &=-|B|^2 w+2h_{1,11}h_{2,12}\lan e_{11,22},A\ran+2h_{1,22}h_{2,21}\lan e_{21,12},A\ran\\ &=-|B|^2 w+4\mu_1\mu_2\lan e_{11,22},A\ran \endaligned \end{equation} where $A=\ep_1\w\cdots\w\ep_n$ and the last step follows from $e_{11,22}=-e_{21,12}=\nu_1\w\nu_2\w e_3\w\cdots\w e_n$. By (\ref{dw}), $$\aligned \n_{e_1}w&=h_{1,11}\lan e_{11},A\ran+h_{2,12}\lan e_{22},A\ran=\mu_1\lan e_{11},A\ran+\mu_2\lan e_{22},A\ran\\ \n_{e_2}w&=h_{1,22}\lan e_{21},A\ran+h_{2,21}\lan e_{12},A\ran=-\mu_1\lan e_{21},A\ran+\mu_2\lan e_{12},A\ran \endaligned$$ and $\n_{e_i}w=0$ for every $i\geq 3$. Hence \begin{equation}\label{dw2} \aligned |\n w|^2=&\sum_i |\n_{e_i}w|^2=\big(\mu_1\lan e_{11},A\ran+\mu_2\lan e_{22},A\ran\big)^2+\big(-\mu_1\lan e_{21},A\ran+\mu_2\lan e_{12},A\ran\big)^2\\ =&\big(\mu_1\lan e_{11},A\ran-\mu_2\lan e_{22},A\ran\big)^2+\big( \mu_1\lan e_{21},A\ran+\mu_2\lan e_{12},A\ran\big)^2\\ &+4\mu_1\mu_2\big(\lan e_{11},A\ran\lan e_{22},A\ran-\lan e_{21},A\ran\lan e_{12},A\ran\big) \endaligned \end{equation} By Lemma 3.2 of \cite{j-x-y3}, \begin{equation}\label{eq1} \lan e_1\w\cdots\w e_n,A\ran\lan e_{11,22},A\ran-\lan e_{11},A\ran\lan e_{22},A\ran+\lan e_{12},A\ran\lan e_{21},A\ran=0. \end{equation} In conjunction with (\ref{La2}), (\ref{dw2}) and (\ref{eq1}), we have \begin{equation}\aligned \De \log w&=w^{-2}(w\De w-|\n w|^2)\\ &=-|B|^2-w^{-2}\Big[\big(\mu_1\lan e_{11},A\ran-\mu_2\lan e_{22},A\ran\big)^2+\big( \mu_1\lan e_{21},A\ran+\mu_2\lan e_{12},A\ran\big)^2\Big] \endaligned \end{equation} whenever $w>0$. We thus have the following results from our previous paper \begin{pro}\cite{j-x-y3} Let $M$ be a minimal submanifold of $\R^{n+m}$ with $G-rank\le 2$ and $w>0$. Then, \begin{equation}\label{dew} \De \log w\leq -|B|^2. \end{equation} \end{pro} \begin{defi}\label{G-conformall} Let $M$ be an $n$-dimensional minimal submanifold in $\ir{n+m}$ , ($m\geq 2$). A point $p\in M$ is called a \textit{G-conformal point}, if there exists an orthonormal basis $\{e_1,\cdots,e_n\}$ of $T_p M$, such that $B_{e_i e_j}=0$ whenever $i\geq 3$ or $j\geq 3$, and $$\lan B_{e_1 e_1},B_{e_1 e_2}\ran=0,\qquad |B_{e_1 e_1}|=|B_{e_1 e_2}|.$$ Moreover if each point of $M$ is a G-conformal point, we call $M$ a \textit{totally G-conformal minimal submanifold}. \end{defi} The formula (\ref{b4}) is for minimal submanifolds in $\ir{n+m}$ with codimension $m\ge 2$. In the present situation we can derive it directly and analyze its accuracy. A straightforward calculation shows (see \cite{Si} \cite{x}) \begin{equation}\label{b} \n^2 B=-\td{\mc{B}}-\ul{\mc{B}}. \end{equation} Here $\n^2$ denotes the trace-Laplace operator acting on any cross-section of a vector bundle over $M$, \begin{equation} \td{\mc{B}}:=B\circ B^t\circ B \end{equation} with $B^t$ denoting the conjugate map of $B$, and \begin{equation} \ul{\mc{B}}:=\sum_{\a=1}^m \big(B_{A^\a A^\a(X),Y}+B_{X,A^\a A^\a(Y)}-2B_{A^\a(X),A^\a(Y)}\big). \end{equation} Hence \begin{equation}\label{b1} \aligned \lan \td{\mc{B}},B\ran&=\lan B\circ B^t\circ B,B\ran=\lan B^t\circ B,B^t\circ B\ran\\ &=\lan B_{e_ie_j},B_{e_ke_l}\ran\lan B_{e_i e_j},B_{e_k e_l}\ran=h_{\a,ij}h_{\a,kl}h_{\be,ij}h_{\be,kl}\\ &=h_{\a,ij}h_{\be,ji}h_{\a,kl}h_{\be,lk}=(A^\a A^\be)_{ii}(A^\a A^\be)_{kk}\\ &=\sum_{\a,\be}\big[\tr(A^\a A^\be)\big]^2=4\mu_1^4+4\mu_2^4 \endaligned \end{equation} where the last step follows from (\ref{sf}), and \begin{equation}\label{b2} \aligned \lan \ul{\mc{B}},B\ran=&\lan B_{A^\a A^\a(e_i),e_j}+B_{e_i,A^\a A^\a(e_j)}-2B_{A^\a(e_i),A^\a(e_j)},\nu_\be\ran \lan B_{e_i,e_j},\nu_\be\ran\\ =&\lan A^\be A^\a A^\a(e_i),e_j\ran \lan A^\be(e_j),e_i\ran+\lan A^\be A^\a A^\a(e_j),e_i\ran\lan A^\be(e_i),e_j\ran\\ &-2\lan A^\be A^\a(e_i),A^\a(e_j)\ran \lan A^\be(e_j),e_i\ran\\ =&(A^\be A^\a A^\a)_{ij}(A^\be)_{ji}+(A^\be A^\a A^\a)_{ji}(A^\be)_{ij}-2(A^\a A^\be A^\a)_{ij}(A^\be)_{ji}\\ =&2\tr(A^\be A^\a A^\a A^\be-A^\a A^\be A^\a A^\be)=2\tr\big([A^\be,A^\a]A^\a A^\be\big)\\ =&\tr\big([A^\be,A^\a]A^\a A^\be\big)+\tr\big([A^\a,A^\be]A^\be A^\a\big)=-\sum_{\a,\be}\tr\big([A^\a,A^\be]^2\big)\\ =&-2\tr\big([A^1,A^2]^2\big)=16\mu_1^2\mu_2^2 \endaligned \end{equation} where \begin{equation} [A^1,A^2]=\left(\begin{array}{cccc} 0 & 2\mu_1\mu_2 & &\\ -2\mu_1\mu_2 & 0 & &\\ & & &\\ & & \multicolumn{1}{c}{\raisebox{0.5ex}[0pt]{\Huge o}}& \end{array}\right). \end{equation} Substituting (\ref{b1}) and (\ref{b2}) into (\ref{b}) gives \begin{equation}\label{b3} \aligned -\f{\lan \n^2 B,B\ran}{|B|^4}&=\f{\lan \td{\mc{B}}+\ul{\mc{B}},B\ran}{|B|^4} =\f{4\mu_1^4+4\mu_2^4+16\mu_1^2\mu_2^2}{(2\mu_1^2+2\mu_2^2)^2}\\ &=1+\f{2\mu_1^2\mu_2^2}{(\mu_1^2+\mu_2^2)^2}\leq \f{3}{2} \endaligned \end{equation} where the equality holds if and only if $\mu_1=\mu_2$. \begin{pro}\label{simons} Let $M$ be an $n$-dimensional minimal submanifold in $\ir{n+m}$ with codimension $m\geq 2$. Then \begin{equation*} \De |B|^2\geq 2|\n B|^2-3|B|^4. \end{equation*} In the case of $G-rank\le 2$ the equality holds at $p\in M$ if and only if $p$ is a G-conformal point. \end{pro} In order to make use of the formula (\ref{b4}), we also need to estimate $|\n B|^2$ in terms of $|\n |B||^2$. Schoen-Simon-Yau \cite{s-s-y} obtained such an estimate for the hypersurface case. It was generalized to arbitrary codimension in \cite{x-y1} and refined and generalized in \cite{x2}. In particular, if $G-rank\le 2$ for $M$, we have a more precise estimate. \begin{pro}\label{kato1} If $M$ is an $n$-dimensional minimal submanifold in $\R^{n+m}$ with $G-rank\le 2$, then \begin{equation}\label{kato} |\n B|^2\geq 2\big|\n |B|\big|^2. \end{equation} The equality holds at $p\in M$, if and only if there exist an orthonormal basis $\{e_1,\cdots,e_n\}$ of $T_p M$ and $\la_1,\la_2\in \R$, such that $B_{e_ie_j}=0$ whenever $i\geq 3$ or $j\geq 3$, $\lan B_{e_1 e_1},B_{e_1 e_2}\ran=0$, $(\n_{e_k}B)_{e_i e_j}=0$ whenever $i\geq 3$, $j\geq 3$ or $k\geq 3$, and \begin{equation}\label{kato2} \aligned (\n_{e_1}B)_{e_1 e_1}&=\la_1 B_{e_1 e_1}-\la_2 B_{e_1 e_2},\\ (\n_{e_2}B)_{e_1 e_1}&=\la_2 B_{e_1 e_1}+\la_1 B_{e_1 e_2}. \endaligned \end{equation} In particular, if $n=2$ and $m=1$, $|\n B|^2= 2\big|\n |B|\big|^2$ holds everywhere. \end{pro} \begin{proof} It is sufficient for us to prove the inequality at the points where $|B|^2\neq 0$. With the same notation $A^\a,\mu_1,\mu_2$ as in (\ref{sf}), the triangle inequality yields \begin{equation}\label{tri}\big|\n |B|^2\big|=\Big|\sum_\a \n |A^\a|^2\Big|\leq \sum_\a \big|\n |A^\a|^2\big|. \end{equation} By the Schwarz inequality, we obtain \begin{equation}\aligned\label{na1} \big|\n |B|\big|^2&=\f{\big|\n |B|^2\big|^2}{4|B|^2}\leq\f{\Big(\sum_\a \big|\n |A^\a|^2\big|\Big)^2}{4\sum_\a |A^\a|^2} = \f{\Big(\sum_\a \f{\nA{\a}}{|A^\a|}\cdot |A^\a|\Big)^2}{4\sum_\a |A^\a|^2}\\ &\leq \f{\sum\Big(\f{{\nA{\a}}^2}{|A^\a|^2}\Big)\cdot \sum_\a |A^\a|^2}{4\sum_\a |A^\a|^2} =\sum_\a \f{{\nA{\a}}^2}{4 |A^\a|^2}\\ &=\f{{\nA{1}}^2}{4 |A^1|^2}+\f{{\nA{2}}^2}{4 |A^2|^2}. \endaligned \end{equation} Note that here and in the sequel we set $ \f{{\nA{\a}}^2}{4 |A^\a|^2}=0$ whenever $|A^\a|=0$. Since $|A^\a|^2=\sum_{i,j}h_{\a,ij}^2$, \begin{equation}\label{na} \n_{e_k}|A^\a|^2=2h_{\a,ij}h_{\a,ijk} \end{equation} with \begin{equation} h_{\a,ijk}:=\lan (\n_{e_k}B)_{e_i e_j},\nu_\a\ran. \end{equation} As shown above, the assumption $G-rank\le 2$ implies the existence of a local orthonormal tangent frame field $\{e_i\}$ on an open domain $U$ as shown before, such that $B_{e_i e_j}\equiv 0$ whenever $i\geq 3$ or $j\geq 3$. Hence for arbitrary $i,j\geq 3$, $$0=\n_{e_k}(B_{e_i e_j})=(\n_{e_k}B)_{e_i e_j}+B_{\n_{e_k}e_i,e_j}+B_{e_i,\n_{e_k}e_j}=(\n_{e_k}B)_{e_i e_j}$$ holds for all $k$, i.e. \begin{equation}\label{b6} h_{\a,ijk}=0\qquad \forall i,j\geq 3. \end{equation} It immediately follows that \begin{equation}\label{b5} 0=\lan \n_{e_k} H,\nu_\a\ran=\sum_i h_{\a,iik}=h_{\a,11k}+h_{\a,22k}. \end{equation} In conjunction with (\ref{sf}), (\ref{na}) and (\ref{b5}), we get $$\aligned \nA{1}^2&=4\sum_k\big(\sum_{i,j}h_{1,ij}h_{1,ijk}\big)^2=4\sum_k(h_{1,11}h_{1,11k}+h_{1,22}h_{1,22k})^2\\ &=16\mu_1^2\sum_k h_{1,11k}^2=8|A^1|^2\sum_k h_{1,11k}^2 \endaligned$$ and moreover \begin{equation}\label{na2} \f{\nA{1}^2}{|A^1|^2}=8\sum_k h_{1,11k}^2. \end{equation} A similar calculation shows \begin{equation}\label{na3} \f{\nA{2}^2}{|A^2|^2}=8\sum_k h_{2,12k}^2. \end{equation} Substituting (\ref{na2}) and (\ref{na3}) into (\ref{na1}) implies \begin{equation}\label{na4} \big|\n|B|\big|^2\leq 2\sum_{k}h_{1,11k}^2+2\sum_k h_{2,12k}^2. \end{equation} On the other hand, \begin{equation}\aligned\label{na5} |\n B|^2=&\sum_{\a,i,j,k}h_{\a,ijk}^2\geq \sum_{i,j,k}h_{1,ijk}^2+\sum_{i,j,k}h_{2,ijk}^2\\ =&\ (h_{1,111}^2+h_{1,221}^2+h_{1,122}^2+h_{1,212}^2)+(h_{1,112}^2+h_{1,121}^2+h_{1,211}^2+h_{1,222}^2)\\ &+\sum_{k\geq 3}(h_{1,11k}^2+h_{1,1k1}^2+h_{1,k11}^2+h_{1,22k}^2+h_{1,2k2}^2+h_{1,k22}^2)\\ &+(h_{2,121}^2+h_{2,112}^2+h_{2,211}^2+h_{2,222}^2)+(h_{2,122}^2+h_{2,212}^2+h_{2,221}^2+h_{2,111}^2)\\ &+\sum_{k\geq 3}(h_{2,12k}^2+h_{2,k12}^2+h_{2,2k1}^2+h_{2,21k}^2+h_{2,k21}^2+h_{2,1k2}^2)\\ \geq& 4\sum_k h_{1,11k}^2+4\sum_k h_{2,12k}^2. \endaligned \end{equation} Here we have used (\ref{b6}), (\ref{b5}) and $h_{\a,ijk}=h_{\a,ikj}$, which is an immediate corollary of the Codazzi equations. Combining this with (\ref{na4}) and (\ref{na5}) yields (\ref{kato}). Now we determine the conditions ensuring that equality in (\ref{kato}) holds true at $p\in M$. Obviously $|\n B|^2=2\big|\n |B|\big|^2$ requires all the equalities in (\ref{tri}), (\ref{na1}) and (\ref{na5}) hold simultaneously. It is easily seen that equality holds in (\ref{na5}) if and only if $h_{\a,ijk}=0$ whenever one of the indices is no less than $3$. Hence by (\ref{na}), \begin{equation}\aligned \n |A^1|^2=&(2h_{1,11}h_{1,111}+2h_{1,22}h_{1,221})e_1+(2h_{1,11}h_{1,112}+2h_{1,22}h_{1,222})e_2\\ =&4\mu_1(h_{1,111}e_1+h_{1,112}e_2),\\ \n |A^2|^2=&(2h_{2,12}h_{2,121}+2h_{2,21}h_{2,211})e_1+(2h_{2,12}h_{2,122}+2h_{2,21}h_{2,212})e_2\\ =&4\mu_2(h_{2,112}e_1-h_{2,111}e_2), \endaligned \end{equation} and \begin{equation}\aligned v_1:=\f{\n |A^1|^2}{|A^1|}&=2\sqrt{2}(h_{1,111}e_1+h_{1,112}e_2),\\ v_2:=\f{\n |A^2|^2}{|A^2|}&=2\sqrt{2}(h_{2,112}e_1-h_{2,111}e_2). \endaligned \end{equation} (\ref{tri}) and (\ref{na1}) hold true if and only if the following 2 conditions are satisfied: (i) $\n |A^1|^2$ and $\n |A^2|^2$ point in the same direction, (ii) $(|v_1|,|v_2|)$ and $(\mu_1,\mu_2)$ are linearly depedent. Hence there exist $\la_1,\la_2\in \R$, such that $$\aligned h_{1,111}e_1+h_{1,112}e_2&=\mu_1(\la_1 e_1+\la_2 e_2),\\ h_{2,112}e_1-h_{2,111}e_2&=\mu_2(\la_1 e_1+\la_2 e_2). \endaligned$$ This is equivalent to \begin{equation}\aligned (\n_{e_1}B)_{e_1 e_1}&=\mu_1\la_1 \nu_1-\mu_2\la_2 \nu_2=\la_1 B_{e_1 e_1}-\la_2 B_{e_1 e_2},\\ (\n_{e_2}B)_{e_1 e_1}&=\mu_1\la_2 \nu_1+\mu_2\la_1 \nu_2=\la_2 B_{e_1 e_1}+\la_1 B_{e_1 e_2}. \endaligned \end{equation} \end{proof} In conjunction with (\ref{b4}) and (\ref{kato}), we arrive at \begin{equation}\label{si2} \De|B|^2\geq 4\big|\n|B|\big|^2-3|B|^4. \end{equation} \Section{Curvature estimates}{Curvature estimates} We are ready to derive the curvature estimates, in a manner similar to \cite{e-h}. When $w>0$, we put $v:=w^{-1}$, then (\ref{dew}) is equivalent to \begin{equation}\label{dew20} \De v\geq |B|^2 v+v^{-1}|\n v|^2. \end{equation} From (\ref{dew20}) and (\ref{si2}), a straightforward calculation shows $$\aligned &\De\big(|B|^{2s}v^q\big)\\ =&\De\big(|B|^{2s}\big)v^q+|B|^{2s}\De v^q+2\lan \n |B|^{2s},\n v^q\ran\\ \geq&s|B|^{2s-2}\Big(4\big|\n |B|\big|^2-3|B|^4\Big)v^q+4s(s-1)|B|^{2s-2}\big|\n |B|\big|^2 v^q\\ &+q|B|^{2s}v^{q-1}\big(|B|^2 v+v^{-1}|\n v|^2\big)+q(q-1)|B|^{2s}v^{q-2}|\n v|^2\\ &+4sq|B|^{2s-1}v^{q-1}\lan \n|B|,\n v\ran\\ \geq&(-3s+q)|B|^{2s+2}v^q+4s^2|B|^{2s-2}\big|\n |B|\big|^2v^q\\ &+q^2|B|^{2s}v^{q-2}|\n v|^2+4sq|B|^{2s-1}v^{q-1}\lan \n |B|,\n v\ran. \endaligned$$ It follows that \begin{equation}\label{es3} \De\big(|B|^{2s}v^q\big)\geq (-3s+q)|B|^{2s+2}v^q \end{equation} for arbitrary $s,q\geq 1$. Let $t=2s+1$, then \begin{equation} \De\big(|B|^{t-1}v^q\big)\geq \big(q-\f{3t-3}{2}\big)|B|^{t+1}v^q \end{equation} for arbitrary $t\geq 3$ and $q\geq 1$. Whenever $q>\f{3t-3}{2}$, putting $C_1(t,q)=\big(q-\f{3t-3}{2}\big)^{-1}$ gives \begin{equation} |B|^{2t}v^{2q}\eta^{2t}\leq C_1 \De\big(|B|^{t-1}v^q\big)|B|^{t-1}v^q\eta^{2t} \end{equation} with $\eta$ being an arbitrary smooth function in $M$ with compact supporting set. Integrating both sides of the above inequality over $M$ implies $$\aligned &\int_M |B|^{2t}v^{2q}\eta^{2t}*1\\ \leq &C_1\int_M \De\big(|B|^{t-1}v^q\big)|B|^{t-1}v^q\eta^{2t}*1\\ =&-C_1\int_M \Big\lan \n\big(|B|^{t-1}v^q\big),\n\big(|B|^{t-1}v^q \eta^{2t}\big)\Big\ran*1\\ =&-C_1\int_M \Big|\n\big(|B|^{t-1}v^q\big)\Big|^2\eta^{2t}*1-2tC_1\int_M |B|^{t-1}v^q\eta^{2t-1} \lan \n\big(|B|^{t-1}v^q\big),\n \eta\ran*1\\ \leq&-C_1\int_M \Big|\n\big(|B|^{t-1}v^q\big)\Big|^2\eta^{2t}*1+C_1\int_M \Big|\n\big(|B|^{t-1}v^q\big)\Big|^2\eta^{2t}*1\\ &+C_1t^2\int_M |B|^{2t-2}v^{2q}\eta^{2t-2}|\n \eta|^2*1\\ \leq&C_1t^2\Big(\f{t-1}{t}\ep^{\f{t}{t-1}}\int_M |B|^{2t}v^{2q}\eta^{2t}*1+\f{1}{t}\ep^{-t}\int_M v^{2q}|\n \eta|^{2t}*1\Big) \endaligned $$ for arbitrary $\ep>0$. Here we have used Stokes' theorem and Young's inequality. Choosing $\ep$ such that $C_1t(t-1)\ep^{\f{t}{t-1}}=\f{1}{2}$ gives \begin{equation}\label{es1} \Big(\int_M |B|^{2t}v^{2q}\eta^{2t}*1\Big)^{\f{1}{t}}\leq C_2(t,q)\Big(\int_M v^{2q}|\n \eta|^{2t}*1\Big)^{\f{1}{t}} \end{equation} for arbitrary $t\geq 3$ and $q>\f{3t-3}{2}$. \begin{thm}\label{t1} Let $M$ be an $n$-dimensional minimal submanifold (not necessarily complete) in $\R^{n+m}$ with $G-rank\le 2$ and positive $w$-function on $M$. Let $\rho:M\times M\ra \R$ be a distance function on $M$, such that $|\n \rho(\cdot,p)|\leq 1$ for each $p\in M$. Fix $p_0\in M$, and denote by $B_R=B_R(p_0):=\{p\in M:\rho(p,p_0)<R\}$ the distance ball centered at $p_0$ and of radius $R$. Assume $B_{R_0}\subset B_R\subset\subset M$, then for arbitrary $t\geq 3$ and $q>\f{3t-3}{2}$, there exists a positive constant $C_3$, depending only on $t$ and $q$, such that \begin{equation}\label{es2} \big\| |B|^{2}v^{\f{2q}{t}}\big\|_{L^t(B_{R_0})} \leq C_3(R-R_0)^{-2}\big\|v^{\f{2q}{t}}\big\|_{L^t(B_R)}. \end{equation} with $v:=w^{-1}$. \end{thm} \begin{proof} We let $\psi$ be a standard bump function on $[0,\infty)$ with $\text{supp}(\psi)\subset [0,R)$, $\psi\equiv 1$ on $[0,R_0]$ and $|\psi'|\leq c_0(R-R_0)^{-1}$. Inserting $\eta=\psi\circ \rho(\cdot,p_0)$ in (\ref{es1}), we have \begin{equation} \aligned &\big\| |B|^2v^{\f{2q}{t}}\big\|_{L^t(B_{R_0})}=\Big(\int_{B_{R_0}}|B|^{2t}v^{2q}*1\Big)^{\f{1}{t}}\leq \Big(\int_M |B|^{2t}v^{2q}\eta^{2t}*1\Big)^{\f{1}{t}}\\ \leq& C_2\Big(\int_M v^{2q}|\n \eta|^{2t}*1\Big)^{\f{1}{t}} =C_2\Big(\int_{B_R}v^{2q}|\psi'|^{2t}|\n \rho(\cdot,p_0)|^{2t}*1\Big)^{\f{1}{t}}\\ \leq& C_3(R-R_0)^{-2}\Big(\int_{B_R} v^{2q}*1\Big)^{\f{1}{t}} =C_3(R-R_0)^{-2}\big\|v^{\f{2q}{t}}\big\|_{L^t(B_R)}. \endaligned \end{equation} \end{proof} Furthermore, the mean value inequality for subharmonic functions on minimal submanifolds in Euclidean space can be applied to deduce a pointwise estimate for $|B|^2$. \begin{thm}\label{t2} Our assumption of $M$ is the same as in Theorem \ref{t1}. Denote by $D_R=D_R(p_0)$ the exterior ball centered at $p_0$ and of radius $R$, then for every $t\geq 3$, there exists a positive constant $C_4$ only depending on $t$, such that \begin{equation}\label{es4} (|B|^2 v^3)(p_0)\leq C_4 R^{-2}(\max_{D_R}v)^3\Big(\f{V(R)}{V(\f{R}{2})}\Big)^{\f{1}{t}}. \end{equation} Here $V(R)=V(p_0,R):=\text{Vol}(D_R(p_0)).$ \end{thm} \begin{proof} Let $F: M\ra \R^{n+m}$ be the isomorphic immersion and denote by $r:M\times M\ra \R$ the restriction of the Euclidean distance function. Without loss of generality one can assume $F(p_0)=0$ for $p_0\in M$, then $r^2(\cdot,p_0)=\lan F,F\ran.$ This extrinsic distance function $r$ on $M$ satisfies the assumptions of Theorem \ref{t1}. Letting $q=\f{3t}{2}$ in (\ref{es1}) yields \begin{equation}\label{es5} \Big(\int_M |B|^{2t}v^{3t}\eta^{2t}*1\Big)^{\f{1}{t}}\leq C_2\Big(\int_M v^{3t}|\n \eta|^{2t}*1\Big)^{\f{1}{t}}. \end{equation} Let $\eta$ be a cut-off function on $M$ with $\text{supp}\ \eta\subset B_R$, $\eta|_{B_{\f{R}{2}}}\equiv 1$ and $|\n \eta|\leq c_0R^{-1}$ (the construction of the auxiliary function is the same as in Theorem \ref{t1}). Then \begin{equation} \Big(\int_M v^{3t}|\n \eta|^{2t}*1\Big)^{\f{1}{t}}\leq C_5(t)R^{-2}(\max_{D_R}v)^3V(R)^{\f{1}{t}}. \end{equation} By (\ref{es3}), $|B|^{2t}v^{3t}$ is a subharmonic function on $M$, and by the mean value inequality, \begin{equation}\label{es6} \Big(\int_M |B|^{2t}v^{3t}\eta^{2t}*1\Big)^{\f{1}{t}}\geq \Big(\int_{D_{\f{R}{2}}}|B|^{2t}v^{3t}*1\Big)^{\f{1}{t}}\geq (|B|^2 v^3)(p_0)V\left(\f{R}{2}\right)^{\f{1}{t}}. \end{equation} In conjunction with (\ref{es5})-(\ref{es6}) we arrive at (\ref{es4}). \end{proof} From the preceding curvature estimates we immediately get the following Bernstein type theorem. \begin{thm} Let $M$ be an $n$-dimensional complete minimal submanifold in $\R^{n+m}$ with $G-rank\le 2$ and a positive $w$-function. If $M$ has polynomial volume growth and the function $v=w^{-1}$ has growth \begin{equation} \max_{D_R(p_0)}v=o(R^{\f{2}{3}}) \end{equation} for a fixed point $p_0$, then $M$ has to be an affine linear subspace. \end{thm} \begin{rem} Here, we say that $M$ has polynomial volume growth iff there exists $l\geq 0$ with $V(R)=V(p_0,R)=O(R^l)$. \end{rem} \begin{proof} Let $c_1$ be a positive constant such that \begin{equation}\label{growth} V(R)\leq c_1 R^l. \end{equation} Now we claim \begin{equation}\label{claim} \liminf_{k\ra \infty}\f{V(2^{k+1})}{V(2^k)}\leq 2^l. \end{equation} Otherwise, there are $\ep>0$ and a positive integer $N$, such that for any $k\geq N$, $$\f{V(2^{k+1})}{V(2^k)}\geq 2^l+\ep.$$ Thus, $$\f{V(2^k)}{(2^k)^l}\geq \f{V(2^N)(2^l+\ep)^{k-N}}{(2^N)^l(2^l)^{k-N}} =\f{V(2^N)}{(2^N)^l}\Big(\f{2^l+\ep}{2^l}\Big)^{k-N}.$$ It follows that $$\lim_{k\ra \infty} \f{V(2^k)}{(2^k)^l}=+\infty$$ which contradicts (\ref{growth}). (\ref{claim}) implies the existence of a sequence $\{k_i:i\in \Bbb{N}\}$, such that $k_i<k_j$ whenever $i<j$, $\lim_{i\ra \infty}k_i=\infty$ and $$\f{V(2^{k_i+1})}{V(2^{k_i})}\leq 2^l.$$ then putting $R=R_i:=2^{k_i+1}$ and letting $t=3$ in (\ref{es4}) give \begin{equation} (|B|^2 v^3)(p_0)\leq C_4 2^{\f{l}{3}}R_i^{-2}(\max_{D_{R_i}}v)^3 \end{equation} Since $\max_{D_R}v=o(R^{\f{2}{3}})$, letting $i\ra \infty$ yields $|B|^2=0$ at $p_0$. For arbitrary $p\in M$, put $R_0:=r(p,p_0)$, then the triangle inequality implies $D_{R}(p)\subset D_{R+R_0}(p_0)$ for any $R\geq 0$, hence $$\f{V(p,R)}{R^l}\leq \f{V(p_0,R+R_0)}{R^l}\leq \f{c_1(R+R_0)^l}{R^l}$$ which means $V(p,R)=O(R^l)$. Similarly one can show $\max_{D_R(p)}v=o(R^{\f{2}{3}})$ for arbitrary $p$. Thereby one can proceed as above to arrive at $|B|^2=0$ at $p$. Hence $|B|\equiv 0$ on $M$ and $M$ has to be affine linear. \end{proof} \Section{Graphical cases}{Graphical cases} Let $f=(f^1,\cdots,f^n):\Om\subset \R^n\ra \R^m$ be a vector-valued function, then the graph $M=\{(x,f(x)):x\in \Om\}$ is an embedded submanifold in $\R^{n+m}$. Let $\{\ep_i,\ep_{n+\a}\}$ be the standard orthonormal basis, and put $A=\ep_1\w \cdots\w \ep_n$, then as shown in \cite{j-x}, the $w$-function is positive everywhere on $M$ and the volume element of $M$ is \begin{equation} *1=v\ dx^1\w\cdots\w dx^n, \end{equation} where \begin{equation}\label{v-f} v=w^{-1}=\left[\det\Big(\de_{ij}+\sum_\a \f{\p f^\a}{\p x^i}\f{\p f^\a}{\p x^j}\Big)\right]^{\f{1}{2}}. \end{equation} Without loss of generality we can assume $f(0)=0$. Denote $p_0=(0,0)$, then \begin{equation} D_R=D_R(p_0)=\{(x,f(x)):|x|^2+|f(x)|^2\leq R^2\}. \end{equation} Denote \begin{equation} \Om_R=\{x\in \Om:|x|^2+|f(x)|^2\leq R^2\}, \end{equation} then obviously $\Om_R\subset \Bbb{D}^n(R)$ and $D_R$ is just the graph over $\Om_R$, where $\Bbb{D}^n(R)$ is the $n$-dimensional Euclidean ball of radius $R$. Hence if \begin{equation} \max_{D_R}v\leq CR^l, \end{equation} then \begin{equation}\aligned V(R)&=\int_{D_R}*1=\int_{\Om_R}v dx^1\w\cdots\w dx^n\\ &\leq \max_{D_R}v\cdot \text{Vol}(\Om_R)\leq CR^l\text{Vol}(\Bbb{D}^n(R))\\ &=C\om_n R^{n+l} \endaligned \end{equation} with $\om_n$ being the volume of the $n$-dimensional unit Euclidean ball. This means that the exterior balls of a graph have polynomial volume growth whenever the $v$-function has polynomial growth. This fact leads us to the following result. \begin{thm}\label{t3} Let $M=\{(x,f(x)):x\in \R^n\}$ be an entire minimal graph given by a vector-valued function $f:\R^n\ra \R^m$ with $G-rank\le 2$. If the slope of $f$ satisfies \begin{equation} \De_f=\left[\det\Big(\de_{ij}+\sum_{\a}\f{\p f^\a}{\p x^i}\f{\p f^\a}{\p x^j}\Big)\right]^{\f{1}{2}}=o(R^{\f{2}{3}}), \end{equation} where $R^2=|x|^2+|f(x)|^2$, then $f$ has to be an affine linear function. \end{thm} Now we study $2$-dimensional cases. It is well-known that every oriented $2$-dimensional Riemannian manifold $M$ admits a local isothermal coordinate chart around any point. More precisely, each $p\in M$ has a coordinate neighborhood $(U;u,v)$, such that $$g=\la^2(du^2+dv^2)$$ on $U$ with a positive function $\la$. In fact, for minimal entire graphs, one can find a global isothermal coordinate chart: \begin{lem}(\cite{o} \S 5)\label{iso} Let $M=\{(x,f(x):x\in \R^2\}$ be a $2$-dimensional entire minimal graph in $\R^{2+m}$, then there exists a nonsigular linear transformation \begin{equation}\aligned u_1&=x_1\\ u_2&=ax_1+bx_2,\qquad (b>0) \endaligned \end{equation} such that $(u_1,u_2)$ are global isothermal parameters for $M$. \end{lem} Equipped with this tool, we can obtain another Bernstein type theorem for entire minimal graphs of dimension $2$. \begin{thm}\label{t4} Let $f:\R^2\ra \R^m$ $(x^1,x^2)\mapsto (f^1,\cdots,f^m)$ be an entire solution of the minimal surface equations \begin{equation}\label{mi} \Big(1+\Big|\f{\p f}{\p x^2}\Big|^2\Big)\f{\p^2 f}{(\p x^1)^2}-2\Big\lan \pd{f}{x^1},\pd{f}{x^2}\Big\ran\f{\p^2 f}{\p x^1\p x^2}+\Big(1+\Big|\pd{f}{x^1}\Big|^2\Big)\f{\p^2 f}{(\p x^2)^2}=0. \end{equation} If for some $\ep>0$, \begin{equation} \De_f=\det\Big(\de_{ij}+\sum_\a \pd{f^\a}{x^i}\pd{f^\a}{x^j}\Big)^{\f{1}{2}}=O(R^{1-\ep}) \end{equation} with $R=|x|$, then $f$ has to be affine linear. \end{thm} \begin{proof} By Lemma \ref{iso}, one can find a global isothermal coordinate $(u_1,u_2)$ for the entire minimal graph $M:=\{(x,f(x)):x\in \R^2\}$, i.e. \begin{equation}\aligned g&=\la^2\big((du^1)^2+(du^2)^2\big)=\la^2\big((dx^1)^2+(a\ dx^1+b\ dx^2)^2\big)\\ &=\la^2\big((1+a^2)(dx^1)^2+2ab\ dx^1 dx^2+b^2 (dx^2)^2\big). \endaligned \end{equation} In other words, the metric is given by \begin{equation}\label{metric2} (g_{ij})=\la^2\left(\begin{array}{cc} 1+a^2 & ab\\ ab & b^2 \end{array}\right). \end{equation} Denote the two eigenvalues of $\left(\begin{array}{cc} 1+a^2 & ab\\ ab & b^2 \end{array}\right)$ by $\la_1^2\geq \la_2^2>0$, then \begin{equation}\label{v1} v=\det(g_{ij})^{\f{1}{2}}=\la^2\la_1\la_2. \end{equation} Since $M$ is a graph, any function $\varphi$ on $M$ can be regarded as a function on $\R^2$. Denote \begin{equation} \p_i \varphi=\pd{\varphi}{x^i},\qquad D\varphi=(\p_1 \varphi,\p_2 \varphi) \end{equation} and let $\n \varphi$ be the gradient vector of $\varphi$ on $M$ with respect to $g$. Since the largest eigenvalue of $(g^{ij})$ equals the multiplicative inverse of the smallest eigenvalue of $(g_{ij})$, which is $\la^{-2}\la_2^{-2}$, we have \begin{equation*} |\n \varphi|^2=g^{ij}\p_i\varphi\p_j\varphi\leq \la^{-2}\la_2^{-2}|D\varphi|^2 \end{equation*} i.e. \begin{equation} |\n \varphi|\leq \la^{-1}\la_2^{-1}|D\varphi|=\Big(\f{\la_1}{\la_2}\Big)^{\f{1}{2}}v^{-\f{1}{2}}|D\varphi|. \end{equation} Given $0<R_0<R$, let $\psi$ be a standard bump function, such that $\text{supp}\ \psi\subset [0,R)$, $\psi\equiv 1$ on $[0,R_0]$ and $|\psi'|\leq c_0(R-R_0)^{-1}$. Taking $\eta(x,f(x))=\psi(|x|)$ in (\ref{es1}) gives \begin{equation}\aligned &\Big(\int_{\Bbb{D}^2(R_0)}|B|^{2t}v^{2q+1}dx^1 dx^2\Big)^{\f{1}{t}}\leq \Big(\int_M |B|^{2t}v^{2q}\eta^{2t}*1\Big)^{\f{1}{t}}\\ \leq&C_2\Big(\int_M v^{2q}|\n\eta|^{2t}*1\Big)^{\f{1}{t}}=C_2\Big(\int_M v^{2q}\Big(\f{\la_1}{\la_2}\Big)^t v^{-t}|D\eta|^{2t}*1\Big)^{\f{1}{t}}\\ \leq&C_6(R-R_0)^{-2}\Big(\int_{\Bbb{D}^2(R)}v^{2q-t+1}dx_1dx_2\Big)^{\f{1}{t}}\\ \leq& C_6(R-R_0)^{-2}(\max_{\Bbb{D}^2(R)} v)^{\f{2q+1}{t}-1}(\pi R^2)^{\f{1}{t}}\\ =&C_7\Big(1-\f{R_0}{R}\Big)^{-2}R^{-2+\f{2}{t}}(\max_{\Bbb{D}^2(R)} v)^{\f{2q+1}{t}-1} \endaligned \end{equation} with $C_6$ and $C_7$ being positive constants depending only on $t,q,a$ and $b$. Letting $q=\f{3t-1}{2}$ gives $\f{2q+1}{t}-1=2$. Thus the growth condition of $v$ implies \begin{equation} \Big(\int_{\Bbb{D}^2(R_0)}|B|^{2t}v^{3t}dx^1 dx^2\Big)^{\f{1}{t}}\leq C_8\Big(1-\f{R_0}{R}\Big)^{-2}R^{\f{2}{t}-2\ep}. \end{equation} Taking $t=\f{2}{\ep}$ and then letting $R\ra +\infty$ force $|B|(x,f(x))=0$ whenever $|x|<R_0$. Finally by letting $R_0\ra +\infty$ we get the Bernstein type result. \end{proof} Given a vector-valued function $f:\R^2\ra \R^m$, denote by $$Df=Df(x):=\Big(\pd{f^\a}{x^i}\Big)$$ the Jacobi matrix of $f$ at $x\in \R^2$. $Df$ can also be seen as a linear mapping from $\R^2$ to $\R^m$. Obviously $Df (Df)^T$ is a nonnegative definite symmetric matrix, whose engenvalues are denoted by $\mu_1^2\geq \mu_2^2\geq 0$. It is easy to check that $\mu_1$ and $\mu_2$ are just the critical values of the function $$v\in \R^2\backslash 0\mapsto \f{\big|(Df)(v)\big|}{|v|}$$ and for any bounded domain $\mc{D}\subset \R^2$, $$\mu_1\mu_2=\f{\text{Area}\big(Df(\mc{D})\big)}{\text{Area}(\mc{D})}.$$ In matrix terminology, $\mu_1^2\mu_2^2$ equals the squared sum of all the $2\times 2$-minors of $Df$, i.e. \begin{equation} \mu_1^2\mu_2^2=\sum_{\a<\be}\Big(\pd{f^\a}{x^1}\pd{f^\be}{x^2}-\pd{f^\a}{x^2}\pd{f^\be}{x^1}\Big)^2. \end{equation} When $m=2$, $\mu_1\mu_2$ then is the absolute value of $J_f:=\det(Df)$. As shown in (\ref{v-f}), the metric matrix of the graph given by $f$ is \begin{equation} (g_{ij})=I_2+Df (Df)^T. \end{equation} Thus the two eigenvalues of $(g_{ij})$ are $1+\mu_1^2$ and $1+\mu_2^2$, and \begin{equation} v^2=\det(g_{ij})=(1+\mu_1^2)(1+\mu_2^2). \end{equation} Now we additionally assume that $f$ is an entire solution of the minimal surface equations. Then as shown in (\ref{metric2}), there exists a positive function $\la$ on $M$ and two positive constants $\la_1,\la_2$, depending only on $a$ and $b$, such that \begin{equation} 1+\mu_1^2=\la^2\la_1^2\qquad 1+\mu_2^2=\la^2\la_2^2. \end{equation} Hence \begin{equation} \aligned \mu_1^2\mu_2^2&=(\la^2\la_1^2-1)(\la^2\la_2^2-1)=\la_1^2\la_2^2\la^4-(\la_1^2+\la_2^2)\la^2+1\\ &=v^2-\f{\la_1^2+\la_2^2}{\la_1\la_2}v+1. \endaligned \end{equation} Note that $\f{\la_1^2+\la_2^2}{\la_1\la_2}$ is a constant. Once $v$ has polynomial growth, $\mu_1\mu_2$ also has polynomial growth of the same order, and vice versa. Therefore one can obtain an equivalent form of Theorem \ref{t4} as follows. \begin{thm} Let $f:\R^2\ra \R^m$ $(x^1,x^2)\mapsto (f^1,\cdots,f^m)$ be an entire solution of the minimal surface equations. If for some $\ep>0$, \begin{equation}\label{Ja1} \sum_{\a<\be}\Big(\pd{f^\a}{x^1}\pd{f^\be}{x^2}-\pd{f^\a}{x^2}\pd{f^\be}{x^1}\Big)^2=O(R^{2(1-\ep)}) \end{equation} with $R=|x|$, then $f$ has to be affine linear. If $m=2$, the condition (\ref{Ja1}) is equivalent to \begin{equation} |J_f|:=|\det(Df)|=O(R^{1-\ep}). \end{equation} \end{thm} Similarly we have a version of Theorem \ref{t3} for the minimal surface case. \begin{thm} Let $f:\R^2\ra \R^m$ $(x^1,x^2)\mapsto (f^1,\cdots,f^m)$ be an entire solution of the minimal surface equations. If \begin{equation}\label{Ja2} \sum_{\a<\be}\Big(\pd{f^\a}{x^1}\pd{f^\be}{x^2}-\pd{f^\a}{x^2}\pd{f^\be}{x^1}\Big)^2=o(R^{\f{4}{3}}) \end{equation} with $R^2=|x|^2+|f(x)|^2$, then $f$ has to be affine linear. If $m=2$, the condition (\ref{Ja2}) is equivalent to \begin{equation} |J_f|:=|\det(Df)|=o(R^{\f{2}{3}}). \end{equation} \end{thm} \begin{rem} Obviously, the above result is also a generalization of that of \cite{h-s-v1}. \end{rem} \Section{Discussions}{Discussions} We wish to discuss the case of a minimal surface $M$ in $\R^{2+m}$. It is natural to ask under which conditions the equality in (\ref{b4}), (\ref{dew}) or (\ref{kato}) holds. With the aid of Lemma \ref{iso}, one can get a sufficient condition for equality in (\ref{dew}). \begin{pro} If $M$ is a $2$-dimensional entire minimal graph in $\R^{2+m}$, then \begin{equation}\label{dew1} \De \log w=-|B|^2. \end{equation} \end{pro} \begin{proof} By Lemma \ref{iso}, there exists a nonsingular linear transformation \begin{equation}\label{iso2} \left(\begin{array}{c} u^1 \\ u^2\end{array}\right)= \left(\begin{array}{cc} 1 & 0\\ a & b\end{array}\right) \left(\begin{array}{c} x^1 \\ x^2\end{array}\right) \end{equation} such that $(u^1,u^2)$ are global isothermal parameters for $M$, where $a$ and $b>0$ are constants. Hence there is a positive function $\la$ on $M$, such that the metric $g$ on $M$ can be expressed as \begin{equation}g=\la^2\big((du^1)^2+(du^2)^2\big).\end{equation} As shown in (\ref{v1}), $$w^{-1}=v=\la^2\la_1\la_2$$ with $\la_1^2\geq \la_2^2>0$ being eigenvalues of $\left(\begin{array}{cc} 1+a^2 & ab\\ ab & b^2 \end{array}\right)$. Thus \begin{equation*} \log w=-\log (\la^2)-\log(\la_1\la_2) \end{equation*} and moreover \begin{equation}\label{dew2} \De \log w=-\De \log (\la^2). \end{equation} The Gauss curvature $K$ of $M$ is given by (see e.g. \cite{j}) \begin{equation}\label{dew3} K=-\f{1}{2}\De \log(\la^2). \end{equation} On the other hand, let $\{e_1,e_2\}$ be an orthonormal basis of $T_p M$, with $p$ an arbitrary point in $M$. Since $M$ is minimal, $B_{e_1e_1}+B_{e_2e_2}=0$ and the Gauss equation yields \begin{equation}\label{dew4} K=\mathrm{det} B_{e_ie_j}=-\f{1}{2}|B|^2. \end{equation} Finally combining (\ref{dew2}), (\ref{dew3}), (\ref{dew4}) yields (\ref{dew1}). \end{proof} Let $M$ be a Riemann surface and $F=(F^1,\cdots,F^{n+m}):M\ra \R^{2+m}$ be an isomorphic immersion. Every $p\in M$ has a coordinate neighborhood $(U;u,v)$ such that $g=\la^2(du^2+dv^2)$ on $U$. Now we introduce the complex coordinate \begin{equation*} w=u+\sqrt{-1}v. \end{equation*} It is well-known that $F$ is minimal if and only if all components of $F$ are harmonic functions on $M$, i.e. $\pd{F}{w}$ is a vector-valued holomorphic function on $U$; here and in the sequel \begin{equation}\label{hol}\aligned \f{\p}{\p w}=\f{1}{2}\Big(\f{\p}{\p u}-\sqrt{-1}\f{\p}{\p v}\Big),\qquad &\pf{\bar{w}}=\f{1}{2}\Big(\pf{u}+\sqrt{-1}\pf{v}\Big),\\ dw=du+\sqrt{-1}dv,\qquad &d\bar{w}=du-\sqrt{-1}dv. \endaligned \end{equation} While $\pd{F}{w}$ depends on the choice of local coordinate, the vector-valued holomorphic $1$-form $$\p F:=\pd{F}{w}dw$$ is independent of these local coordinates and can be well-defined on the whole surface, where $dw=du+\sqrt{-1}dv$. Similarly we can define $\bar{\partial} F:=\pd{F}{\bar{w}}d\bar{w}.$ With the symmetric bi-linear form $$\lan (a_1,\cdots,a_N),(b_1,\cdots,b_N)\ran=\sum_{i=1}^N a_i b_i,$$ since $(u,v)$ are isothermal parameters, it is well known and easy to check that \begin{equation}\label{w1} \Big\lan \pd{F}{w},\pd{F}{w}\Big\ran=0,\qquad \Big\lan \pd{F}{w},\pd{F}{\bar{w}}\Big\ran>0 \end{equation} which is equivalent to \begin{equation} \lan \p F,\p F\ran=0,\qquad \lan \p F,\bar{\partial} F\ran>0. \end{equation} Similarly, one can define \begin{equation} \p^2 F:=\ppd{F}{w}dw^2,\qquad \bar{\partial}^2 F:=\ppd{F}{\bar{w}}d\bar{w}^2. \end{equation} Then the minimality of $F$ implies that $\p^2 F$ is a vector-valued holomorphic $2$-form. \begin{pro}\label{simons2} For a fixed point $p$ in a minimal surface $M\subset \R^{2+m}$, the following statements are equivalent: (a) $\De |B|^2=|\n B|^2-3|B|^4$ at $p$; (b) $p$ is a G-conformal point; (c) $\big\lan B_{\pf{w}\pf{w}},B_{\pf{w}\pf{w}}\big\ran=0$ at $p$, where $w$ is a local complex coordinate near $p$; (d) $\lan \p^2 F,\p^2 F\ran=0$ at $p$. \end{pro} \begin{proof} The equivalence of (a) and (b) has been proved in Proposition \ref{simons}. Since $(u,v)$ is an isothermal coordinate, $\pf{u}$ and $\pf{v}$ have the same length and are orthogonal to each other, hence $p$ is an holomorphic-like point if and only if \begin{equation} |B_{uu}|=|B_{uv}|,\qquad \lan B_{uu},B_{uv}\ran=0. \end{equation} Here and in the sequel, $B_{uu}:=B_{\pf{u}\pf{u}}$, $B_{uv}:=B_{\pf{u}\pf{v}}$ and so on. By using (\ref{hol}) one can get \begin{equation}\label{b11} B_{ww}=\f{1}{2}B_{uu}-\f{\sqrt{-1}}{2}B_{uv}. \end{equation} It implies \begin{equation}\aligned \lan B_{ww},B_{ww}\ran=\f{1}{4}\big(|B_{uu}|^2-|B_{uv}|^2\big)-\f{\sqrt{-1}}{2}\lan B_{uu},B_{uv}\ran \endaligned \end{equation} and hence (b) and (c) are equivalent. Since $$(T_p M)\otimes \Bbb{C}=\text{span}\Big\{\pd{F}{w},\pd{F}{\bar{w}}\Big\}$$ there exist two complex numbers $\mu_1$ and $\mu_2$, such that \begin{equation}\label{w2} \n_{\pf{w}}\pf{w}=\Big(\ppd{F}{w}\Big)^T=\mu_1 \pd{F}{w}+\mu_2\pd{F}{\bar{w}}. \end{equation} By (\ref{w1}), \begin{equation}\aligned 0&=\f{1}{2}\pf{w}\Big\lan \pd{F}{w},\pd{F}{w}\Big\ran=\Big\lan \ppd{F}{w},\pd{F}{w}\Big\ran\\ &=\Big\lan \mu_1\pd{F}{w}+\mu_2\pd{F}{\bar{w}},\pd{F}{w}\Big\ran=\mu_2\Big\lan \pd{F}{\bar{w}},\pd{F}{w}\Big\ran. \endaligned \end{equation} Hence $\mu_2=0$ and moreover \begin{equation} \aligned \Big\lan \ppd{F}{w},\ppd{F}{w}\Big\ran&=\Big\lan \Big(\ppd{F}{w}\Big)^N,\Big(\ppd{F}{w}\Big)^N\Big\ran+\Big\lan \Big(\ppd{F}{w}\Big)^T,\Big(\ppd{F}{w}\Big)^T\Big\ran\\ &=\lan B_{ww},B_{ww}\ran+\mu_1^2\Big\lan \pd{F}{w},\pd{F}{w}\Big\ran=\lan B_{ww},B_{ww}\ran. \endaligned \end{equation} Thus (c) is equivalent to (d). \end{proof} Define \begin{equation} \om:=\lan \p^2 F,\p^2 F\ran=\Big\lan \ppd{F}{w},\ppd{F}{w}\Big\ran dw^4 \end{equation} then it is easy to check that the definition of $\om$ is independent of the choice of coordinate, and $$\pf{\bar{w}}\Big\lan \ppd{F}{w},\ppd{F}{w}\Big\ran=2\Big\lan \pf{w}\Big(\f{\p^2 F}{\p w\p \bar{w}}\Big),\ppd{F}{w}\Big\ran=0$$ implies $\om$ is a homolomorphic $4$-form on $M$. By using Proposition \ref{simons2} we immediately get the following corollary. \begin{cor} Let $M$ be a minimal surface in $\ir{2+m}$, then $M$ is totally G-conformal if and only if the holomorphic $4$-form $\om:=\lan \p^2 F,\p^2 F\ran$ vanishes everywhere. \end{cor} \begin{cor} Let $M=\{(x,f(x)):x\in\R^2\}$ be an entire minimal graph in $\R^4$. Then $M$ is totally G-conformal if and only if at least one of the following 3 cases occurs: (i) $f:\R^2\ra \R^2$ is a holomorphic function; (ii) $f$ is anti-holomorphic; (iii) $f$ is affine linear. \end{cor} \begin{proof} Let $(u_1,u_2)$ be the global isothermal parameters on $M$ given in (\ref{iso2}). Denote $z:=u_1+\sqrt{-1}u_2$ and \begin{equation} \phi_i=\pd{x^i}{z},\qquad \phi_{2+\a}=\pd{f^\a}{z}. \end{equation} then $\pd{F}{z}=(\phi_1,\phi_2,\phi_3,\phi_4)$ and (\ref{w1}) yields $\phi_1^2+\phi_2^2+\phi_3^2+\phi_4^2=0$. By (\ref{iso2}), $\phi_1$ and $\phi_2$ are both constants, denote \begin{equation} d:=\phi_1^2+\phi_2^2, \end{equation} then \begin{equation} \phi_3^2+\phi_4^2=-(\phi_1^2+\phi_2^2)=-d. \end{equation} If $d=0$, then $\phi_4=\pm \sqrt{-1}\phi_3$ and hence \begin{equation} \ppd{F}{z}=(\phi'_1,\phi'_2,\phi'_3,\phi'_4)=(0,0,\phi'_3,\pm \sqrt{-1}\phi'_3). \end{equation} It follows that \begin{equation} \Big\lan \ppd{F}{z},\ppd{F}{z}\Big\ran=(\phi'_3)^2-(\phi'_3)^2=0 \end{equation} and $M$ is totally holomorphic-like. As show in \cite{h-s-v}, $d=0$ implies $f$ is holomorphic or anti-holomorphic, and vice versa. If $d\neq 0$, then \begin{equation}\label{hol1} -d=\phi_3^2+\phi_4^2=(\phi_3+\sqrt{-1}\phi_4)(\phi_3-\sqrt{-1}\phi_4) \end{equation} implies $\phi_3-\sqrt{-1}\phi_4$ is an entire function having no zeros, hence there is an entire function $H(z)$, such that \begin{equation} \phi_3-\sqrt{-1}\phi_4=e^{H(z)}. \end{equation} Substituting it into (\ref{hol1}) gives \begin{equation} \phi_3+\sqrt{-1}\phi_4=-de^{-H(z)}. \end{equation} In conjunction with the above two equations we have \begin{equation} \phi_3=\f{1}{2}(e^H-d e^{-H}),\qquad \phi_4=\f{\sqrt{-1}}{2}(e^H+d e^{-H}). \end{equation} Thus \begin{equation} \aligned \Big\lan \ppd{F}{z},\ppd{F}{z}\Big\ran&=(\phi'_1)^2+(\phi'_2)^2+(\phi'_3)^2+(\phi'_4)^2\\ &=\f{1}{4}(e^H+d e^{-H})^2(H')^2-\f{1}{4}(e^H-d e^{-H})^2(H')^2\\ &=d(H')^2 \endaligned \end{equation} which is identically zero if and only if $H$ is a constant function. In this case, $\phi_i$ and $\phi_{2+\a}$ are all constants on $M$, hence $M$ has to be an affine plane. \end{proof} For the sequel, we put \begin{equation} (\n B)_{uuv}:=(\n_{\pf{v}}B)\Big(\pf{u},\pf{u}\Big),\qquad (\n B)_{www}:=(\n_{\pf{w}}B)\Big(\pf{w},\pf{w}\Big) \end{equation} and so on. Then (\ref{kato2}) says that there are $\xi_1,\xi_2\in \R$, such that \begin{equation}\label{nb10} \aligned (\n B)_{uuu}&=\xi_1 B_{uu}-\xi_2 B_{uv}\\ (\n B)_{uuv}&=\xi_2 B_{uu}+\xi_1 B_{uv}. \endaligned \end{equation} \begin{pro}\label{p10} For a fixed point $p$ in a minimal surface $M\subset \R^{2+m}$, the following statements are equivalent: (a) $|\n B|^2=2\big|\n |B|\big|^2$ at $p$; (b) There is an isothermal coordinate chart $(U; u,v)$ around $p$, such that $(\n B)_{www}=\ze B_{ww}$ at $p$, with $w=u+\sqrt{-1}v$ and $\ze\in \Bbb{C}$; (c) For an arbitrary isothermal coordinate chart $(U; u,v)$ around $p$, there is $\ze\in \Bbb{C}$, such that $(\n B)_{www}=\ze B_{ww}$ at $p$, with $w=u+\sqrt{-1}v$. \end{pro} \begin{proof} The equivalence of (b) and (c) is obvious, so it is sufficient to prove the equivalence of (a) and (b). Similarly to Section \ref{s1}, one can choose an isothermal coordinate neighborhood $(U;u,v)$ of $p$, such that $$\lan B_{uu}, B_{uv}\ran=0\qquad \text{at }p.$$ Then by Proposition \ref{kato1}, (a) is equivalent to (\ref{nb10}). By (\ref{hol}), one can obtain \begin{equation}\label{nb11} (\n B)_{www}=\f{1}{2}(\n B)_{uuu}-\f{\sqrt{-1}}{2}(\n B)_{uuv} \end{equation} with the aid of the Codazzi equations. If (\ref{nb10}) holds, letting $\ze:=\xi_1-\sqrt{-1}\xi_2$ and combining with (\ref{b11}) and (\ref{nb11}) implies \begin{equation} \aligned \ze B_{ww} &=\f{1}{2}(\xi_1 B_{uu}-\xi_2 B_{uv})-\f{\sqrt{-1}}{2}(\xi_1 B_{uv}+\xi_2 B_{uu})\\ &=\f{1}{2}(\n B)_{uuu}-\f{\sqrt{-1}}{2}(\n B)_{uuv}=(\n B)_{www}. \endaligned \end{equation} Conversely, if $(\n B)_{www}=\ze B_{ww}$, then by letting $\xi_1=\text{Re}\ze$ and $\xi_2=-\text{Im}\ze$, one can proceed similarly to above to get (\ref{nb10}). Therefore (a) and (b) are equivalent. \end{proof} \begin{cor} Let $M$ be a totally G-conformal minimal surface in $\R^4$, then \begin{equation}\label{nb12} |\n B|^2=2\big|\n |B|\big|^2 \end{equation} holds at any $p\in M$ satisfying $|B|^2(p)>0$. \end{cor} \begin{proof} Since $M$ is totally holomorphic-like, $B_{uu}$ and $B_{uv}$ have the same length and are orthogonal to each other. Since $\dim N_pM=2$ and $|B|^2(p)>0$ we conclude that $N_p M=\text{span}\{B_{uu},B_{uv}\}$ and moreover $$N_p M\otimes \C=\text{span}\{B_{ww},B_{\bar{w}\bar{w}}\}.$$ Thus there are $\mu_3,\mu_4\in \Bbb{C}$, such that $$(\n B)_{www}=\mu_3 B_{ww}+\mu_4 B_{\bar{w}\bar{w}}.$$ Differentiating both sides of $\lan B_{ww},B_{ww}\ran=0$ yields \begin{equation}\aligned 0&=\f{1}{2}\pf{w}\lan B_{ww},B_{ww}\ran=\lan \n_{\pf{w}}(B_{ww}),B_{ww}\ran\\\ &=\lan (\n B)_{www}+2B_{\n_{\pf{w}}\pf{w},\pf{w}},B_{ww}\ran\\ &=\lan (\n B)_{www},B_{ww}\ran+2\mu_1\lan B_{ww},B_{ww}\ran\\ &=(\mu_3+2\mu_1)\lan B_{ww} ,B_{ww}\ran+\mu_4\lan B_{\bar{w}\bar{w}},B_{ww}\ran\\ &=\mu_4|B_{ww}|^2 \endaligned \end{equation} where we have used (\ref{w2}). Hence $\mu_4=0$ and then (\ref{nb12}) follows from Proposition \ref{p10}. \end{proof} \bibliographystyle{amsplain}
\section{Introduction} Soft-hadronic observables measured in the relativistic heavy-ion experiments are well reproduced by perfect-fluid hydrodynamics or by viscous hydrodynamics with a small viscosity to entropy ratio \cite{Chaudhuri:2006jd,Dusling:2007gi,Luzum:2008cw,Song:2007fn,Bozek:2009dw,Schenke:2010rr}. Nevertheless, the use of such approaches at the very early stages of the collisions encounters conceptual difficulties. Thermalization times shorter than a fraction of a fermi (used in the perfect-fluid approaches) cannot be explained within microscopic models of the collisions. On the other hand, viscous hydrodynamics is based on an implicit assumption that one can make an expansion around an isotropic background. If the shear correction is large, a new framework incorporating large momentum-space anisotropies into the leading order of the approximation may be useful. Such a new approach has been introduced in \cite{Florkowski:2010cf,Martinez:2010sc,Ryblewski:2010bs,Martinez:2010sd,Ryblewski:2011aq,Martinez:2012tu,Ryblewski:2012rr} and we refer to it below as to the {\it anisotropic hydrodynamics}. \section{Anisotropic hydrodynamics} Anisotropic hydrodynamics is based on the equations \begin{eqnarray} \partial_\mu T^{\mu \nu} &=& 0, \label{enmomcon} \\ \partial_\mu \sigma^{\mu} &=& \Sigma, \label{engrow} \end{eqnarray} which express the energy-momentum conservation and entropy production laws. The energy-momentum tensor $T^{\mu \nu}$ has the structure \begin{eqnarray} T^{\mu \nu} = \left( \varepsilon + P_{\perp}\right) U^{\mu}U^{\nu} - P_{\perp} \, g^{\mu\nu} - (P_{\perp} - P_{\parallel}) V^{\mu}V^{\nu}, \label{Taniso} \end{eqnarray} where $P_{\parallel}$ is the longitudinal pressure and $P_{\perp}$ is the transverse pressure. In the limit $P_{\parallel}=P_{\perp}=P$, Eq.~(\ref{Taniso}) reproduces the energy-momentum tensor of the perfect fluid. Similarly, the entropy production law (\ref{engrow}) is reduced to the entropy conservation law, if we take $\Sigma=0$. The four-vector $U^{\mu}$ describes the flow of matter, while $V^{\mu}$ defines the beam ($z$) axis. In the general case, we use the parameterizations $U^\mu = (u_0 \cosh \vartheta, u_x, u_y, u_0 \sinh \vartheta)$ and $V^\mu = ( \sinh \vartheta, 0, 0, \cosh \vartheta)$, where $u_x$ and $u_y$ are the transverse components of the four-velocity field and $\vartheta$ is the longitudinal fluid rapidity. The entropy flux $\sigma^{\mu}$ equals $\sigma \, U^\mu$, where $\sigma$ is the non-equilibrium entropy density. One can show \cite{Florkowski:2010cf} that instead of $P_{\parallel}$ and $P_{\perp}$ it is more convenient to use the entropy density $\sigma$ and the {\it anisotropy parameter} $x$ as two independent variables (one may use the approximation $P_{\parallel}/P_{\perp} \approx x^{-3/4}$). Similarly to standard hydrodynamics with vanishing baryon chemical potential, the energy density $\varepsilon$, the entropy density $\sigma$, and the anisotropy parameter $x$ are related through the \textit{generalized} equation of state \cite{Ryblewski:2010bs} \begin{eqnarray} \varepsilon (x,\sigma)&=& \varepsilon_{\rm qgp}(\sigma) r(x), \label{epsilon2b} \\ \nonumber P_\perp (x,\sigma)&=& P_{\rm qgp}(\sigma) \left[r(x) + 3 x r^\prime(x) \right], \label{PT2b} \\ \nonumber P_\parallel (x,\sigma)&=& P_{\rm qgp}(\sigma) \left[r(x) - 6 x r^\prime(x) \right]. \label{PL2b} \end{eqnarray} where the functions $\varepsilon_{\rm qgp}$ and $P_{\rm qgp}$ define the realistic QCD equation of state constructed in Ref. \cite{Chojnacki:2007jc}. The function $r(x)$ characterizes properties of the fluid which exhibits the pressure anisotropy $x$ \cite{Florkowski:2010cf} \begin{equation} r(x) = \frac{ x^{-\frac{1}{3}}}{2} \left[ 1 + \frac{x \arctan\sqrt{x-1}}{\sqrt{x-1}}\right]. \label{RB} \end{equation} In the isotropic case $x = 1$, $r(1)=1$, $r^\prime(1)=0$, and Eq.~(\ref{epsilon2b}) is reduced to the standard equation of state used in \cite{Chojnacki:2007jc}. The function $\Sigma$ in Eq.~(\ref{engrow}) defines the entropy source. We use the form proposed in \cite{Florkowski:2010cf} \begin{equation} \Sigma(\sigma,x) = \frac{(1-\sqrt{x})^{2}}{\sqrt{x}}\frac{\sigma}{\tau_{\rm eq}}, \label{entropys} \end{equation} where the time-scale parameter $\tau_{\rm eq}$ controls the rate of equilibration. In this work we use the constant value $\tau_{\rm eq}$ = 1 fm. In Refs. \cite{Martinez:2010sc,Martinez:2010sd,Martinez:2012tu} the medium dependent $\tau_{\rm eq}$ was used, which was inversely proportional to the typical transverse momentum scale in the system. If a constant value of $\tau_{\rm eq}$ is used, the system approaches the perfect fluid behavior for $\tau \gg \tau_{\rm eq}$. In the limit of small anisotropy Eq.~(\ref{entropys}) is consistent with the quadratic form of the entropy production in the Israel-Stewart theory \cite{Martinez:2010sc}. Far from equilibrium, hints for the form of $\Sigma$ are lacking, although we may expect some suggestions from the AdS/CFT correspondence \cite{Heller:2011ju}. Thus, for large anisotropies the formula (\ref{entropys}) should be treated as an assumption defining the dynamics of the system. \section{Initial conditions and freeze-out} In the general 3+1 dimensional case we have to solve Eqs. (\ref{enmomcon}) and (\ref{engrow}) for $\sigma$, $x$, $u_x$, $u_y$, and $\vartheta$, which depend on $\tau,{\bf x}_\perp=(x,y)$, and $\eta$ ($\tau$ is the proper time and $\eta$ is the space-time rapidity). We fix the initial starting time to $\tau_{\rm 0} =0.25$ fm. Similarly to other hydrodynamic calculations, we assume that there is no initial transverse flow, $u_x(\tau_{\rm 0},{\bf x}_\perp,\eta) = u_y(\tau_0,{\bf x}_\perp,\eta) = 0$ and that the initial longitudinal rapidity of the fluid is equal to space-time rapidity, $\vartheta(\tau_0,{\bf x}_\perp,\eta) = \eta$. In this text we present the results for two scenarios: i) the initial source is strongly {\it oblate} in momentum space, $x(\tau_0,{\bf x}_\perp,\eta) =100$, and ii) the source is {\it prolate} in momentum space, $x(\tau_0,{\bf x}_\perp,\eta) = 0.032$; the latter value is chosen since $r(100) = r(0.032)$. The initial entropy density profile has the form \begin{equation} \sigma_0(\eta,{\bf x}_\perp) = \sigma(\tau_0,\eta,{\bf x}_\perp) = \varepsilon_{\rm gqp}^{-1} \left[ \varepsilon_{\rm i} \, \tilde{\rho}(b,\eta,{\bf x}_\perp) \right], \label{sig2} \end{equation} where $b$ is the impact parameter, and $\tilde{\rho}(b,\eta,{\bf x}_\perp)$ is the normalized density of sources, $\tilde{\rho}(b,\eta,{\bf x}_\perp) = \rho(b,\eta,{\bf x}_\perp)/\rho(0,0,0)$, for details see \cite{Ryblewski:2012rr}. The quantity $\varepsilon_{\rm i}$ is the initial energy density at the center of the system created in the most central collisions. Its value is fixed by the measured multiplicity, separately for two different physical scenarios considered in this paper. We use $\varepsilon_{\rm i}$ = 48.8 GeV/fm$^3$ and 80.1 GeV/fm$^3$ for $x_0 = 100$ and $x_0 = 0.032$, respectively. \begin{figure}[t] \begin{center} \includegraphics[angle=0,width=0.5\textwidth]{piK_v2_cen3.eps} \caption{\small (Color online) Transverse-momentum dependence of the elliptic flow coefficient $v_2$ of $\pi^{+}+K^{+}$ calculated for $c=20-40$\% ($b=7.84$ fm) at midrapidity and for $\tau_{\rm eq}= 1.0$ fm/c; $x_0=100$ (dashed blue lines) and $x_0=0.032$ (dotted green lines). The results are compared to the PHENIX Collaboration data (red dots) \cite{Adler:2003kt}. } \end{center} \label{fig1} \end{figure} \begin{figure}[t] \begin{center} \includegraphics[angle=0,width=0.5\textwidth]{v1_vs_eta_mid3.eps} \caption{\small (Color online) Pseudorapidity dependence of the directed flow of charged particles for the centrality bin $c=5-40$\% and $\tau_{\rm eq}=1.0$ fm/c; $x_0=100$ (dashed blue lines) and $x_0=0.032$ (dotted green lines). The results are compared to the experimental data from STAR (red dots) \cite{Abelev:2008jga} and PHOBOS (green squares) \cite{Back:2005pc}. } \end{center} \label{fig2} \end{figure} The evolution is determined by the hydrodynamic equations until the entropy density drops to $\sigma_{\rm f} = 1.79$ fm$^{-3}$, which for $x=1$ corresponds to the temperature $T_{\rm f} = 150$ MeV. According to the single-freeze-out scenario, at this moment the abundances and momenta of particles are expected to be fixed. The processes of particle production and decays of unstable resonances are described by using {\tt THERMINATOR} \cite{Kisiel:2005hn,Chojnacki:2011hb}, which applies the Cooper-Frye formalism to generate hadrons on the freeze-out hypersurface. \section{Results} The model results describing the pseudorapidity distributions, transverse-momentum spectra, the elliptic and directed flow coefficients, and the HBT radii have been obtained with different initial anisotropies of pressure; $x_0=100$ (dashed blue lines) and $x_0=0.032$ (dotted green lines). In all of the considered cases we find good agreement between the model results and the data. Moreover, we find that the results obtained with different initial anisotropies are practically the same. This is so because we have adjusted the initial energy density separately for two different values of $x_0$. A larger (smaller) initial energy density is used for the initially prolate (oblate) system. Our results describing the elliptic and directed flow are shown and compared to the RHIC data (Au+Au collisions at the highest beam energy $\sqrt{s_{\rm NN}}$ = 200 GeV) in Figs.~1 and 2. Our results indicate that the final hadronic observables are not sensitive to the early anisotropy of pressure. The flows are built up during the whole time evolution of the system, hence the relatively short early anisotropic stage does not influence the results. In our opinion, the insensitivity of the hadronic observables helps us to circumvent the early thermalization/isotropization puzzle. \medskip {\bf Acknowledgements:} This work was supported by the Polish Ministry of Science and Higher Education under Grant No.~N N202 263438 and the United States National Science Foundation under Grant No. PHY-1068765.
\section{Introduction} Consider a system of stochastic equations of the form \begin{equation} \label{Eq:TheSystem} \begin{cases} X_t^\epsilon & = x_0 + \int_0^t f(X_s^\epsilon,Y_s^\epsilon) ds + \int_0^t g(X_s^\epsilon,Y_s^\epsilon) dW_s\\ Y_t^\epsilon & = y_0 + \epsilon^{-1} \int_0^t b(X_s^\epsilon,Y_s^\epsilon) ds + \epsilon^{-1/2} \int_0^t \sigma (X_s^\epsilon, Y_s^\epsilon) d\tilde{W}_s, \end{cases} \end{equation} where $X_t^\epsilon$ is a $d_x$-dimensional process, $Y_t^\epsilon$ a $d_y$-dimensional process, $W$ and $\tilde{W}$ are two independent Brownian motions of dimensions $d_{x}$ and $d_{y}$, and the functions $b,\sigma, f \text{ and } g$ have the right dimensions. This type of system models the dynamics of two sets of interacting variables evolving in different time scales. The difference between time scales is controlled by the parameter $\epsilon$. In many domains the most interesting case of study is that of the regime when $\epsilon \ll 1$, i.e. the situation in which $X^\epsilon$ is evolving very slowly compared to $Y^\epsilon$ (for this reason we will frequently denominate them as \emph{slow scale} and \emph{fast scale} variables respectively). This regime may be studied by singular perturbation techniques as the ones presented in \cite{bensoussan_asymptotic_1978} for deterministic models: instead of looking at the system with a small $\epsilon$, we study the limit of \eqref{Eq:TheSystem} as $\epsilon \rightarrow 0$ (when it exists) and estimate the error induced by this approximation. There exist several analytical works with applications in different domains on the described type of approximation for stochastic models. For example in \cite{majda_mathematical_2001} a climate model is considered and studied on the advection scale (i.e. in the time scale of the slow variable). In \cite{fouque_derivatives_2000} and \cite{fouque_singular_2003} a system similar to \eqref{Eq:TheSystem} is presented and studied for pricing derivatives in the context of stochastic volatility models. A complete study with rather general hypothesis on the coefficients of the system is found in \cite{pardoux_poisson_2001} and \cite{pardoux_poisson_2003}. In these papers a system with a fast scale ergodic diffusion is considered. More precisely if \begin{equation} \label{Eq:ErgodicDiffusion} Y_t^x = y_0 + \int_0^t b(x,Y_s^x) ds + \int_0^t \sigma (x, Y_s^x) d\tilde{W}_s, \end{equation} is ergodic with unique invariant measure $\mu^x$, we might define the \emph{effective equation} \begin{equation} \label{Eq:EffectiveEquation} X_{t} = x_0 + \int_0^t F (X_{s}) ds + \int_0^t G (X_{s}) dW_{s}, \end{equation} with coefficients given by \begin{equation} F(x) = \int f(x,y) \mu^x(dy)\quad G(x) = \sqrt{H(x)}\quad H(x) = \int h(x,y) \mu^x(dy), \label{Eq:DefinitionAverageF} \end{equation} where $h(x,y) = gg^*(x,y)$, and $G(x)$ could be in principle any matrix with square given by $H$, but we choose it to represent the Cholesky decomposition of the positive semi-definite matrix $H$. It follows that under appropriate assumptions $X^{\epsilon} \tendslaw X$ as $\epsilon \rightarrow 0$ (c.f. \cite{pardoux_poisson_2003}). The idea behind this kind of singular perturbation method is that when the difference between scales is large enough, the dynamics of the system behave as if the slow scale would be frozen and the ergodic limit of the fast diffusion would be attained. However, except for a few particular examples, it is not an easy task to find explicit expressions for the averages \eqref{Eq:DefinitionAverageF}. Naturally, this leads to the question of designing numerical methods of approximation of the effective equation. Several methods have been developed for a purely deterministic case (see for example the review \cite{e_heterogeneous_2007}). Most of them are based on choosing a macro-solver for the slow scale in which some information from the fast scale is added via parameters introduction to guarantee the correct approximation. The literature with respect to numerical approximation of the general stochastic case is, to our knowledge, much more restricted. In \cite{e_analysis_2005} the authors present an algorithm based on the use of an approximation scheme for the slow scale (for example the Euler scheme) and at each step of the slow scale another scheme is used to solve for the fast scale contribution; the weak and strong error induced by the scheme is analyzed when considering the particular case of an ODE with random coefficients slow scale equation and a stochastic ergodic fast scale variable (i.e. when $g(x,y)=0$ in \eqref{Eq:TheSystem}). In our work we use a similar approach. We focus on approaching numerically equation \eqref{Eq:EffectiveEquation}. With this objective in mind, we propose a \emph{Multi-scale Decreasing Step (MsDS)} algorithm defined as a composition of an Euler scheme for the slow scale, the decreasing Euler step algorithm and estimator proposed in \cite{lamberton_recursive_2002} for the ergodic average approximation at each step, and a Cholesky decomposition for finding the volatility coefficient. In order to control the total error approximation of this proposed algorithm we need to take into account four effects. First, we need an estimate on the ergodic average approximation at each step. We show that this control is based on the existence, regularity and control of the solution of the Poisson equation associated to the fast scale diffusion \begin{equation} \label{Eq:PoissonEquation} \mathcal{L}^x_y \phi_\psi(x,y) = \psi(x,y) \end{equation} where \begin{equation} \label{Eq:DiffusionOperator} \mathcal{L}^x_y :=\frac{1}{2} \sum\limits_{i,j = 1}^{d_y} a_{ij}(x,y) \frac{\partial^2}{\partial y_i \partial y_j} + \sum\limits_{i=1}^{d_y} b_i(x,y) \frac{\partial}{\partial y_i} \end{equation} with $a:=\sigma\sigma^*$, when considering as sources (i.e. the right hand side functions) the coefficients $F$ and $H$ centered with respect to their respective invariant measures. Second, we need to control the error obtained after performing a Cholesky decomposition. Then, we have to account for discretization errors. Finally, we need to control the error propagation which will be possible under some growth control on the coefficients of the effective equation. The MsDS algorithm strongly converges to the exact solution and proves to be more efficient than a simple Euler scheme for highly oscillating problems. Moreover, it features a non-standard C.L.T. property in the sense that the normalized error distribution converges towards the solution of an SDE. The coefficients appearing in this normalized error SDE depend on the solution of the previously mentioned Poisson problem and are, in general, unknown. Nevertheless, the available explicit expression for them is valuable for the estimation of confidence intervals and eases the task of parameter tuning for actual implementation of the algorithm. We study as well an \emph{extrapolated MsDS (EMsDS)} version of the algorithm, differing from the original one in that it uses a Richardson-Romberg extrapolation of the decreasing step estimator (i.e. a well chosen linear combination of the decreasing step Euler estimator with appropriate parameters) to approach the ergodic averages. As the MsDS, the EMsDS also features a non-standard C.L.T. property and shares the same rate of convergence. However, the extrapolated version has lower asymptotic complexity and hence higher asymptotic efficiency than the original one. \subsection{Outline of the paper} The organization of the paper is as follows: in section \ref{Sec:Algorithm}, we describe the algorithm and state the standing hypothesis and our main results (strong convergence, limit distribution). The proof of the main theorem is presented in section \ref{Sec:MsDSalgorithm} after having reminded some regularity properties of the effective equation and available results on the decreasing Euler estimation algorithm in section \ref{Sec:Preliminaries}. We extend the main results to an extrapolated version of the algorithm that we introduce and study in section \ref{Sec:EMSDSalgorithm}. Finally, we perform some numerical studies in section \ref{Sec:numresults}. The paper ends with an appendix containing the proof of a couple of technical results. \section{The MsDS algorithm} \label{Sec:Algorithm} Let $(\Omega,\ensuremath \mathcal{F},\prob)$ be a probability space and $W$ be an $\ensuremath \mathcal{F}-$adapted Brownian motion. Suppose we are given an independent probability space $(\tilde{\Omega},\tilde{\ensuremath \mathcal{F}},\tilde{\prob})$ and a family of independent Brownian motions $\tilde{W}^{q}, q\in \mathds{Q}$ with an associated filtration $\tilde{\ensuremath \mathcal{F}}^{q}_t:=\sigma\{ \tilde{W}^{q}_s, s\leq t \}$. Define the extended space $(\bar{\Omega},\bar{\ensuremath \mathcal{F}},\bar{\ensuremath \mathcal{F}}_t, \bar{\prob})$ by \begin{align*} \bar{\Omega} &:= \Omega \times \tilde{\Omega}, & \bar{\prob}(d\omega,d\bar{\omega}) &= \prob(d\omega) \tilde{\prob}(d\bar{\omega}), \\ \bar{\ensuremath \mathcal{F}}&:=\ensuremath \mathcal{F} \otimes \tilde{\ensuremath \mathcal{F}}, &\tilde{\ensuremath \mathcal{F}}^q_t&:= \bigvee\limits_{q\in \mathds{Q}; q \leq t} \tilde{F}^q_\infty, &\bar{F}_t &:=\ensuremath \mathcal{F}_t \vee \tilde{\ensuremath \mathcal{F}}^q_t. && \end{align*} Such extended space will be useful for treating independently the noise coming from the Brownian in the effective diffusion and the one related to the approximation of the ergodic diffusion averages. Consider the decreasing step Euler algorithm introduced in \cite{lamberton_recursive_2002} to approach the invariant measure of a recursive diffusion. Let $\{\gamma_k\}_{k\in \ensuremath \mathds{N}}$ be a decreasing sequence of steps satisfying \begin{HypSteps}[\emph{On the sequence of steps for the average estimation algorithm}]\quad \begin{enumerate}[i)] \item $\gamma_k>0$ for all $k$; \item $\gamma_k$ is a sequence of decreasing steps with $\lim\limits_{n\rightarrow\infty} \gamma_k = 0$; \label{Enum:HipSteps_Decreasing} \item $\lim\limits_{k\rightarrow\infty}\Gamma_k = \infty; \text{ where }\Gamma_k:=\sum_{j=0}^k \gamma_j$; \label{Enum:HipSteps_DivergentSum} \item $\sum_{k=1}^{\infty} \left( \frac{\gamma_k^2}{\Gamma_k} \right) < + \infty$. \end{enumerate} \end{HypSteps} For any $q\in \ensuremath \mathds{Q}$, let $\sqrt{\gamma_{k+1}} U_{k+1}^q := \tilde{W}^q_{\Gamma_{k+1}}-\tilde{W}^q_{\Gamma_{k}}$ so that $U_{k+1}$ is a standard Gaussian vector. Let $y_0 \in \ensuremath \mathds{R}^{d_y}$. We define the \emph{decreasing step Euler approximation of the ergodic diffusion} by \begin{align} \DecrAvg{Y}_{0}^{x,q} & = y_0 \notag\\ \DecrAvg{Y}_{k+1}^{x,q} & =\DecrAvg{Y}_{k}^{x,y_0,q} + \gamma_{k+1} b(x,\DecrAvg{Y}_{k}^{x,q})+\sqrt{\gamma_{k+1}} \sigma(x, \DecrAvg{Y}_k^{x,q})U^q_{k+1}, \label{Eq:MicroAlgo} \end{align} and the \emph{decreasing step average estimator} by \begin{equation} \DecrAvg{F}^{k}(x,q) = \frac{1}{\Gamma_k}\sum\limits_{j=1}^{k} \gamma_j f(x,\DecrAvg{Y}^{x,q}_{j-1}). \label{Eq:AverageEstimator} \end{equation} The idea behind the particular form of estimator \eqref{Eq:AverageEstimator} is to take advantage of the ergodicity of the diffusion: the long term time average approaches the invariant measure of the diffusion. Note that the estimator can also be written recursively as \[\DecrAvg{F}^{0}(x,q) = 0; \quad \DecrAvg{F}^{k}(x,q) = \DecrAvg{F}^{k-1}(x,q) + \frac{\gamma_{k}}{\Gamma_k}\left(f(x,\DecrAvg{Y}_{k-1}^{x,q})-\DecrAvg{F}^{k-1}(x,q)\right).\] Evidently, using the same ergodic average argument, it is also possible to use a uniform step estimator of the type $k^{-1}\sum_{j=1}^{k} \gamma_j f(x,\DecrAvg{Y}^{x,q}_{j-1})$ as studied for example in \cite{Talay_second-order_1990}. The main difference between both estimators appears in the type of error that they generate. The uniform step estimator induces two types of errors coming from the truncation of the series and the fact that the \emph{ergodic limit of the approached sequence is not the ergodic limit of the original diffusion}. In contrast, the decreasing Euler scheme estimator eliminates the asymptotic gap between the invariant law of the continuous equation and that of its discretization (see \cite{lamberton_recursive_2002}). Moreover, the decreasing step method features a kind of ``error expansion´´ (as shown in \cite{lemaire_estimation_2005}) when applied to a certain family of functions. These properties are important to show the limit properties of our algorithm and to deduce the extrapolated version. We should remark that we have chosen to work with a simplified version of the algorithm in \cite{lamberton_recursive_2002}: its more general version allows the use of different sequences for the Euler scheme step and for the weights in the average. With this estimator in hand we can define an Euler scheme to approach our effective diffusion. Assuming a time horizon $T$, for $n\in \ensuremath \mathds{N}^*$ we put $t_k = Tk/n$, so that the Euler scheme will be given by \[\check{X}_{t_{k+1}}^{n} = \check{X}_{t_{k}}^{n} + \DecrAvg{F}^{M(n)}(\check{X}_{t_k}^{n},t_k) \Delta t_{k+1} + \DecrAvg{G}^{M(n)}(\check{X}_{t_k}^{n},t_k) \Delta W_{k+1},\] where $\DecrAvg{F}^M$ is defined in \eqref{Eq:AverageEstimator} and $\DecrAvg{G}^{M}(x,q)$ is defined in two steps: First we find $\DecrAvg{H}^{M}(x,q)$ using the decreasing step algorithm as in \eqref{Eq:AverageEstimator} (recall that $h(x,y)=g^*g\, (x,y)$) and then we perform a Cholesky decomposition on it to find $\DecrAvg{G}^{M}(x,q)=\sqrt{\DecrAvg{H}^{M}(x,q)}$. Note that the number of steps in the decreasing Euler estimator, $M$, is expressed as a function of the number of steps in the Euler scheme for the slow scale $n$. The form of $M(n)$ will be clear from the main theorems. It will be easier to work mathematically with a continuous interpolation of the Euler approximation. Let us denote by $\tbar(n) = \lfloor nt\rfloor/n$.We will usually omit the explicit dependence on $n$ and write $\tbar$ when clear from the context. The continuous Euler approximation is then given by \begin{equation} \label{Eq:continuous_approximated} \ApproxVar{X}^{n}_t = x_0 + \int_0^t \DecrAvg{F}^{M(n)}\left(\ApproxVar{X}^{n}_{\tbar[s]},\tbar[s] \right) ds + \int_0^t \DecrAvg{G}^{M(n)} \left(\ApproxVar{X}^{n}_{\tbar[s]},\tbar[s] \right) dW_s, \end{equation} i.e. a linear interpolation from the discrete Euler scheme. Clearly, at times $t_k$ the continuous Euler coincide with the Euler algorithm. All our results will be derived for the continuous version of the algorithm. \subsection{Standing hypothesis and main result} Let us introduce the assumptions under which our main results follow. \begin{HypSlowScale}[\emph{On the slow-scale coefficients}]\quad \begin{enumerate}[i)] \item Lipschitz in x: There exist constants $K, m$ such that for all $x,x' \in \ensuremath \mathds{R}^{d_x}$ and $y \in \ensuremath \mathds{R}^{d_y}$, \[|f(x,y)-f(x',y)|+|g(x,y)-g(x',y)|\leq K |y|^{m}|x-x'|;\] \label{Enum:HipSlowScale_Lipschitz} \item Regularity: $f, h$ belong to $C_{b,p}^{2,r^y}$ for some $r^y>3$, where the subindex $b,p$ means the derivatives $\partial_x^i\partial_y^j$ for $0\leq i\leq 2$ and $0\leq j \leq r^y-i$ are bounded in $x$ and polynomially bounded in $y$; \label{Enum:HipSlowScale_Growth} \item Degeneracy: Either $h$ is identically zero, or it is uniformly non degenerate, that is, there exists $\lambda^\prime_{-} \in \ensuremath \mathds{R}^+_*$ such that $\lambda^\prime_- I \leq h(x,y)$. \end{enumerate} \end{HypSlowScale} Before giving the standing hypothesis on the fast scale equation, recall that we have defined the matrix $a(x,y)=\sigma\sigma^*(x,y)$. \begin{HypFastScale}[\emph{On the fast-scale coefficients}]\quad \begin{enumerate}[i)] \item $a,b \in C_{b,l}^{2,0}$, i.e. they are continuous and linearly bounded in $y$, and $C^2$ and bounded in $x$. \label{Enum:HipFastScale_Regularity} \item The matrix $a$ is uniformly continuous and uniformly non-degenerate and bounded, i.e. there exist $\lambda_-, \lambda_+ \in \ensuremath \mathds{R}^+_*$ such that \[\lambda_ - I \leq a(x,y) \leq \lambda_+ I; \] \label{Enum:HipFastScale_NonDegeneracy} \vspace{-1\baselineskip} \item $\sup\limits_x b(x,y)\cdot y \leq -c_1|y|^2+c_2, $ for some $c_1\in \ensuremath \mathds{R}^*_+,c_2\in \ensuremath \mathds{R}$ \label{Enum:HipFastScale_Recurrence} \end{enumerate} \end{HypFastScale} The regularity and growth hypothesis contained in $\ensuremath {(\mathcal{H}_{s.s.})}$ are assumed to control the error propagation. The main goal of imposing conditions on the fast scale diffusion is to guarantee the existence of an invariant limit for any possible fixed value of $x$ and a uniform control on its averages. For this reasons they are quite restrictive: note that $\ensuremath {(\mathcal{H}_{f.s.})}$ \eqref{Enum:HipFastScale_Regularity} implies $\sup_x |b(x,y)| = O(|y|)$ and $\ensuremath {(\mathcal{H}_{f.s.})}$ \eqref{Enum:HipFastScale_Recurrence} deduces $\lim_{|y|\rightarrow\infty}\sup_x b(x,y)\cdot y = -\infty$, meaning that the drift has at most linear growth in $y$ and that it is mean reverting uniformly in $x$. In turn, the ellipticity and non-degeneracy assumption $\ensuremath {(\mathcal{H}_{f.s.})}$ \eqref{Enum:HipFastScale_NonDegeneracy} is helpful to deduce the uniqueness of the invariant measure. We are ready to state our main Theorem on the MsDS algorithm. Its proof is found in section \ref{Sec:MsDSalgorithm}. \begin{MyTheorem} \label{Theo:ShortMainDecreasingStep} Let $0<\theta<1$, $\gamma_0 \in \ensuremath \mathds{R}^+$ and $\gamma_k = \gamma_0 k^{-\theta}$. Let $M_1$ be a positive constant. Assume $\ensuremath {(\mathcal{H}_{f.s.})}$ and $\ensuremath {(\mathcal{H}_{s.s.})}$. Define $M(n)$ by \[M(n) = \left\lceil M_1 n^{\frac{1}{1-\theta}} \right\rceil,\] then \begin{enumerate}[i.] \item ODE with random coefficients case $(g(x,y)\equiv 0)$: \begin{enumerate}[a)] \item (Strong convergence). There exists a constant $K$ such that \[\EX{\sup\limits_{0\leq t \leq T}\left|X_t - \ApproxVar{X}_t^{n}\right|^2 } \leq K n^{-2[(1-\theta)\wedge \theta]/(1-\theta)}\] \item \label{Enum:TheoLimitDistODE} (Limit distribution of the error). Assume in addition that $r^y \geq 7$ and $ \theta \geq 1/2$. Then \[n \left(X - \ApproxVar{X}^{n} \right) =: \zeta^{n} \Rightarrow \zeta^\infty\] where $\Rightarrow$ denotes convergence in law and $\zeta^\infty$ is the solution of an SDE stated explicitly on Theorem \ref{Theo:AbstractLimitDistribution}. \end{enumerate} \item Full SDE case \begin{enumerate}[a)] \item (Strong convergence). There exists a constant $K$ such that \[\EX{\sup\limits_{0\leq t \leq T}\left|X_t - \ApproxVar{X}_t^{n}\right|^2 } \leq K n^{-[(1-\theta)\wedge 2\theta]/(1-\theta)}\] \item \label{Enum:TheoLimitDistSDE} (Limit distribution of the error). Assume in addition that $r^y \geq 7$ and $\theta \geq 1/3$. Then \[n^{1/2} \left(X - \ApproxVar{X}^{n} \right) =: \zeta^{n} \Rightarrow \zeta^{\infty},\] where $\zeta^{\infty}$ is the solution of an SDE stated explicitly on Theorem \ref{Theo:AbstractLimitDistribution}. \ref{Theo:AbstractLimitDistribution}. \end{enumerate} \end{enumerate} \end{MyTheorem} Note that we study the mean square error of our approximation algorithm towards the effective equation. We perform this strong error analysis to guarantee that the algorithm will be used for applications demanding to approach functions that depend on the whole trajectory (as in finance). As will be clear from Theorem \ref{Theo:AbstractLimitDistribution}, the SDE defining the limit results both for the fully stochastic and the ODE with random coefficients case are explicitly given in terms of the invariant law of the ergodic diffusion and are consequently unknown. Nevertheless, the key point is that, being explicit, they might be estimated numerically for practical purposes. We have announced an extrapolated version of the algorithm. Given that its proper introduction requires a further understanding of the basic algorithm, we postpone the presentation to section \eqref{Sec:EMSDSalgorithm}. \section{Preliminaries} \label{Sec:Preliminaries} In this section we present the main tools needed to analyze the presented algorithm. Let us start by stating properly the stochastic approximation theorem we mentioned in the introduction and that justifies the relation between the effective equation \eqref{Eq:EffectiveEquation} and the original strongly oscillating system \eqref{Eq:TheSystem}. \begin{MyTheorem}[Theorem 4 in \cite{pardoux_poisson_2003}] \label{Intro:Theo:PardouxLimit} Let $b,\sigma,f,g$ be defined as in \eqref{Eq:TheSystem} and $a=\sigma\sigma^*$. Assume we have a recurrence condition of the type $\lim_{|y|\rightarrow \infty} b(x,y)\cdot y=-\infty$, and that the matrix `$a$' is non-degenerate and uniformly elliptic. Assume that $a,b \in C_b^{2,1+\alpha}$, and that $f, g$ are Lipschitz with respect to the $x$ variable uniformly in $y$ and have at most polynomial growth in $y$ and linear growth in $x$. Then, for any $T>0$, the family of processes $\{X_t^\epsilon, 0\leq t \leq T\}_{0< \epsilon \leq 1}$ is weakly relatively compact in $C([0,T]; \ensuremath \mathds{R}^l)$. Any accumulation point $X$ is a solution of the martingale problem associated with the operator $\bar{\mathcal{L}}$. If moreover, the martingale problem is well posed, then $X^\epsilon \tendslaw X$, where $X$ is the unique (in law) diffusion process with generator $\bar{\mathcal{L}}$. \end{MyTheorem} It is worth mentioning that the actual framework of Pardoux and Vertennikov's statement includes the case in which there is an $\epsilon^{-1}$ order term in the slow variable, which complicates the proof with respect to the framework we present here. Note that under the standing hypothesis, the martingale problem is well posed and $X$ in the theorem is the unique solution to \eqref{Eq:EffectiveEquation}. \subsection{A priori estimates} An important result is related to some a priori estimates valid for general SDEs. Since they are quite standard, we will state the result without giving the details of the proof. \begin{MyProposition} \label{Prop:APriori} Let \begin{equation} \vartheta_{t} = \vartheta_0 + \int_0^t V_1 ( \vartheta_s,s ) ds + \int_0^t V_2 ( \vartheta_s,s) dW_{s}, \label{Eq:GeneralSDEApriori} \end{equation} where $V_1, V_2$ are adapted random functions. \begin{enumerate}[i)] \item \label{Enum:AprioriControl} For all $\alpha\geq 2$, \begin{align*} & \EX{\sup\limits_{ 0 \leq t \leq T} |\vartheta_t|^\alpha} \\ & \leq K_\alpha \EX{|\vartheta_0|^\alpha} + K(\alpha,T) \int_0^T \left( \EX{| V_1( \vartheta_s,s ) |^\alpha} + \EX{| V_2( \vartheta_s,s ) |^\alpha} \right) ds\\ & \leq K_\alpha \EX{|\vartheta_0|^\alpha} + K'(\alpha,T) \left( \sup\limits_{ 0 \leq t \leq T} \EX{ | V_1 (\vartheta_t,t) |^\alpha} + \sup\limits_{ 0 \leq t \leq T} \EX{| V_2 (\vartheta_t, t) |^\alpha } \right). \end{align*} \item \label{Enum:AprioriStoppingTime} Assume that $\forall \alpha \geq 2$, \begin{align*} \EX{|V_1(\vartheta_t,t)|^\alpha} + \EX{|V_2(\vartheta_t,t)|^\alpha} &\leq K\left(1+\EX{|\vartheta_t|}^{\alpha}\right). \end{align*} Then, \begin{enumerate}[a)] \item \label{Enum:LemStoppingTime_PowerBound} For $t\in [0,T]$ and $\alpha \geq 2$, $\EX{|\vartheta_t|^{\alpha}}\leq K(\alpha,T) $ \item \label{Enum:LemStoppingTime_SupPowerBound} For $\alpha \geq 2$, $\EX{\sup\limits_{0\leq s\leq t} |\vartheta_s|^{\alpha}} \leq K(\alpha,T) $ $\displaystyle \probX{\sup\limits_{0 \leq s \leq t}\tau_r \leq t }\leq \frac{K'(\alpha,t)}{r^{\alpha}}.$ \end{enumerate} \end{enumerate} \end{MyProposition} \subsection{Cholesky decomposition} The Cholesky decomposition of a positive definite matrix consists in expressing this matrix as the product of a lower triangular matrix and its conjugate transpose. A stability analysis of this procedure is a key point in our analysis for the SDE case behavior of our algorithm. Recall that we denote by $|\cdot|$ the induced operator norm. Let us denote by $\|\cdot\|_F$ the Frobenius norm. Recall that if $H$ is a $d\times d$ matrix \begin{equation} \label{Eq:NormEquivalence} |H| \leq \|H\|_F \leq \sqrt{d}|H|. \end{equation} \begin{MyTheorem}[Theorem 1.1 in \cite{sun_perturbation_1991}] \label{Theo:ControlCholesky} Let $H$ be a $d \times d$ positive definite matrix with Cholesky factorization $H=GG^*$. If $\Delta H$ is a $d\times d$ symmetrical matrix satisfying $ \displaystyle |H^{-1}| \|\Delta H\|_F < 1/2$, then there is a unique Cholesky factorization $H+\Delta H = (G+\Delta G)(G+\Delta G)^*$ and \begin{equation} \frac{\|\Delta G\|_F}{|G|} \leq \sqrt{2} \frac{\kappa \kappa_2(H)}{1+\sqrt{1-2\kappa_2(H)\kappa}}, \end{equation} where $\kappa = |\Delta H\|_F|H|^{-1}$ and $\kappa_2(H)=|H||H^{-1}|$. \end{MyTheorem} Theorem \ref{Theo:ControlCholesky} gives a control on the sensitivity of the Cholesky procedure. In Lemma \ref{Lem:ErrorCholesky} we study the propagation effect at each stage of the Cholesky factorization to say a little bit more on the particular form of the error. Its proof is available We give the proof in Annex \ref{Sec:AnnexCholesky}. \begin{MyLemma} \label{Lem:ErrorCholesky} Suppose the hypothesis of Theorem \ref{Theo:ControlCholesky} hold. Then, \begin{align*} \Delta G_{i,i} &= \frac{\Delta H_{i,i} - 2 \sum\limits_{k=1}^{i-1}\Delta G_{i,k} G_{i,k}}{2G_{i,i}} + O(|\Delta H|^2),\\ \Delta G_{i,j} &= \frac{\Delta H_{i,j} -G_{i,j}\Delta G_{j,j}-\sum\limits_{k=1}^{j-1} (\Delta G_{j,k} G_{i,k}+\Delta G_{i,k} G_{j,k})}{G_{j,j}} + O(|\Delta H|^2) \end{align*} for $i>j$. \end{MyLemma} Lemma \ref{Lem:ErrorCholesky} gives a first order approximation of the error matrix $\Delta G$ knowing the perturbation matrix $\Delta H$. From this lemma, we can deduce on the regularity of the Cholesky approximation. The following Corollary, follows from the definition of $H$ and Lemma \ref{Lem:ErrorCholesky}. \begin{MyCorollary} \label{Cor:RegularityG} Let $H: \ensuremath \mathds{R}^{d} \rightarrow M^{d \times d}$ be $C_b^2$ and non-degenerate (in the sense given in hypothesis $\ensuremath {(\mathcal{H}_{s.s.})}$). Then $G$ is also $C_b^2$ and non-degenerate. \end{MyCorollary} \subsection{Decreasing step Euler algorithm} In this section we present some control and error expansion results valid for the Decreasing Step Euler algorithm. The results here presented are found in \cite{lamberton_recursive_2002} or in the PhD thesis of \cite{lemaire_estimation_2005}. A first interesting property is that the sequence of estimators defined in \eqref{Eq:AverageEstimator} converges almost surely to the ergodic average for any fixed $x$. \begin{MyProposition} \label{Prop:ConvergenceEstimator} Assume $\ensuremath {(\mathcal{H}_{f.s.})}$, and let $\psi:\ensuremath \mathds{R}^{d_x}\times \ensuremath \mathds{R}^{d_y}\rightarrow \ensuremath \mathds{R}$, and suppose that $\psi(x,y) \leq C(x)(1+|y|^\pi)$. Let $\DecrAvg{\Psi}^M(x,q)$ be defined as in \eqref{Eq:AverageEstimator}. Then, for any $x\in \ensuremath \mathds{R}^{d_x}, q\in\ensuremath \mathds{Q}$, \[\DecrAvg{\Psi}^M(x,q) \tendsas \int \psi(x,y)\mu^x(dy),\text{ as } M\rightarrow \infty\] where $\mu^x$ is the invariant measure of \eqref{Eq:ErgodicDiffusion}. \end{MyProposition} \begin{proof} $\ensuremath {(\mathcal{H}_{f.s.})}$ imply that $V(y):= 1 + |y|^2$ is a uniformly in $x$ function satisfying the hypothesis of Theorem 1 in \cite{lamberton_recursive_2002}, from which the claim follows. \end{proof} We have as well a control on the moments of any order of $\DecrAvg{Y}^{x,q}_k$. \begin{MyProposition} \label{Prop:ApproxFiniteMoments time} Let $\pi>0$ and let $\DecrAvg{Y}_k^{x,q}$ be given by \eqref{Eq:MicroAlgo}. Then there exists a constant $K_\pi$ given only by $\pi$, $\lambda_-$, $\lambda_+$ and $\gamma_0$ such that for all $x\in\ensuremath \mathds{R}^{d_x}$ and $q\in \ensuremath \mathds{Q}$ \[\sup_{i \in \ensuremath \mathds{N}} \EX{ |\DecrAvg{Y}_i^{x,q}|^\pi} < K_\pi .\] Moreover, for every $\pi > 1$, \[ \sup_{M\in \ensuremath \mathds{N}} \left(\frac{1}{\Gamma_M}\sum_{i=1}^{M} \gamma_i |\DecrAvg{Y}_i^{x,q}|^\pi \right) < + \infty\] \end{MyProposition} \begin{proof} By Lemma 2 in \cite{lamberton_recursive_2002} given that $U_k^q$ has moments of any order and $V(y) = |y|^2 +1$ satisfies the needed hypothesis uniformly in $x$, we get that for any $\pi\geq 1$ and $q\in \ensuremath \mathds{Q}$, \[ \sup_{i \in \ensuremath \mathds{N}} \EX{ |\DecrAvg{Y}_i^{x,q}|^{2\pi}} \leq \sup_{i \in \ensuremath \mathds{N}} \EX{ V(\DecrAvg{Y}_i^{x,q})^{\pi}} < K_{\pi}.\] The extension to all $\pi>0$ is straightforward. The second claim follows from Theorem 3 in \cite{lamberton_recursive_2002}. \end{proof} Proposition \ref{Prop:ErrorExpansion} is an adaptation of a result appearing in the PhD thesis \cite{lemaire_estimation_2005}. The proof comes from performing a Taylor expansion and reordering the terms in a proper way. For the statement, we introduce in addition to the sequence $\{\gamma_k\}_{\{k\in \ensuremath \mathds{N}^*\}}$ a new sequence that we denote by $\{\eta_k\}_{\{k\in \ensuremath \mathds{N}^*\}}$ (that may be taken equal to the former). This added flexibility will be useful in the following, in particular to prove Proposition \ref{Prop:ControlDecreasingDifferentStep}. We may interpret Proposition \ref{Prop:ErrorExpansion} as an error expansion result. Indeed if we fix $\eta_k=\gamma_k$ satisfying $\ensuremath {(\mathcal{H}_{\gamma})}$ then we will have an explicit expression for the approximation error of the decreasing Euler algorithm. \begin{MyProposition} \label{Prop:ErrorExpansion} Let $\psi:\ensuremath \mathds{R}^{d_x}\times \ensuremath \mathds{R}^{d_y}\rightarrow \ensuremath \mathds{R}$. Under the assumptions of Proposition \ref{Prop:ConvergenceEstimator}, suppose that for each $x\in \ensuremath \mathds{R}^{d_x}$ there exists $\phi_\psi^x:\ensuremath \mathds{R}^{d_y}\rightarrow \ensuremath \mathds{R}$ solution of the centered Poisson equation \begin{equation} \label{Eq:CenteredPoisson} \mathcal{L}_y^x \phi^x_\psi(y) = \psi (x,y)-\int \psi(x,z)\mu^x(dz) \end{equation} Suppose as well for $r\in \ensuremath \mathds{N}$, $r\geq 2$, that $\phi^x_\psi$ is $C^{r}$ in the $y$-variable uniformly in $x$, and $D^r \phi_\psi$ is Lipschitz in $y$ uniformly in $x$. Let $\gamma_k$ and $\eta_k$ be two decreasing sequences with $\gamma_k \rightarrow 0$, $\eta_k\rightarrow 0$, $\Gamma_k=\sum_{1\leq j\leq k}\gamma_k$, $\mathcal{H}_k=\sum_{1\leq j\leq k}\eta_k$. Let $\DecrAvg{Y}_k^{x,q}$ be defined as in \eqref{Eq:MicroAlgo} (with step sequence $\gamma_k$). Then, \begin{equation*} \sum_{k=1}^{M} \eta_k \left(\psi(x,\DecrAvg{Y}^{x,q}_{k-1}) -\int \psi(x,z)\mu^x(dz) \right) = A^{0}_{\psi,M}-N_{\psi,M} - \sum_{i=2}^{r} A^{i}_{\psi,M} -Z^{r}_{\psi,M} \end{equation*} where \begin{align} A^{0}_{\psi,M} & (x,q):= \sum_{k=1}^{M} \frac{\eta_k}{\gamma_k} \left[ \phi_\psi^x(\DecrAvg{Y}^{x,q}_{k})-\phi_\psi^x(\DecrAvg{Y}^{x,q}_{k-1})\right] \label{Eq:A0M}\\ N_{\psi,M} & (x,q):=\sum\limits_{k=1}^{M} \frac{\eta_k}{\sqrt{\gamma_k}} \langle D_y \phi_\psi^x(\DecrAvg{Y}^{x,q}_{k-1}), \sigma(x,\DecrAvg{Y}^{x,q}_{k-1})U_k^q \rangle \label{Eq:NM}\\ A^{2}_{\psi,M} & (x,q):= \frac{1}{2}\sum\limits_{k=1}^{M}\eta_k\left[ D^2\phi_\psi^x(\DecrAvg{Y}^{x,q}_{k-1})\cdot(\sigma(x,\DecrAvg{Y}^{x,q}_{k-1})U^q_k)^{\otimes 2} \right. \label{Eq:A2M}\\ & \hphantom{ (x):= \frac{1}{2}\sum\limits_{k=1}^{M}\gamma_k } - \left. \textrm{Tr}\left( D^2 \phi_\psi^x(\DecrAvg{Y}^{x,q}_{k-1})(\sigma^*\sigma(x,\DecrAvg{Y}^{x,q}_{k-1})\right)\right]\notag\\ A^{i}_{\psi,M} & (x,q):= \sum\limits_{k=1}^{M} \eta_k\gamma_k^{i/2-1} v_\psi^{i,r}(x,\DecrAvg{Y}^{x,q}_{k-1},U_k^q) \label{Eq:AiM} \end{align} for $i = 3,\dots, r$ with \[v_\psi^{i,r}(x,y,z) = \sum\limits_{j\geq i/2}^{i\wedge r} \binom{j}{i-j}\frac{1}{j!} D_y^j \phi_\psi^x(y) \cdot \left\langle b(x,y)^{\otimes(i-j)}, (\sigma(x,y) z)^{\otimes(2j-i)} \right\rangle;\] and \begin{equation} |Z_{\psi,M}^{r}| (x,q)\leq K \sum\limits_{k=1}^{M} \eta_k\gamma_k^{\frac{r-1}{2}} (1+|\DecrAvg{Y}^{x,q}_{k-1}|^{r+1}) (1 + |U_k^q|)^{r+1}.\label{Eq:Z2M} \end{equation} \end{MyProposition} The average of each expansion term will play an important role in our analysis, so that we will present a special notation for them. Indeed, let \begin{equation} \label{Eq:Definition_v_bar} \begin{split} & \bar{v}_\psi^{i,r} (x,y) := \EX{v_\psi^{i,r}(x,y,U_1^0)} \\ & = \sum_{j\geq i/2}^{i\wedge r} \binom{j}{i-j}\frac{1}{j!}D^j_y\phi_\psi^x(y)\cEX{\left\langle b(x,y)^{\otimes(i-j)}, (\sigma(x,y) U_k^q)^{\otimes(2j-i)}\right\rangle}{\tilde{F}_{\Gamma_{k-1}}} \end{split} \end{equation} \begin{MyRemark} \label{Rmk:ZeroExpectation} Consider $A_{\psi,M}^{2i+1}$ for $i\leq \lfloor (r-1)/2 \rfloor$. As $2j-2i-1$ is odd for any $j$ integer and given the fact that the odd powers of a centered Gaussian are centered we deduce \(\bar{v}_\psi^{2i+1,r} = 0.\) Of course this property transfers to $A_{\psi,M}^{2i+1}$ so that $\EX{A_{\psi,M}^{2i+1}}=0$, implying in turn that the terms with an odd index are centered. \end{MyRemark} Under some additional hypothesis, Proposition \ref{Prop:ErrorExpansion} may be used to obtain an $L_2$ control on the error of the approximation. For the sake of the presentation, let us denote from now on \begin{equation}\label{Eq:DefGammaAlpha}\Gamma_M^{[r]} = \sum_{k=1}^M (\gamma_k)^r.\end{equation} Note we have in particular $\Gamma_M^{[1]}=\Gamma_M$. \begin{MyProposition} \label{Prop:ControlDecreasingDifferentStep} Under the assumptions of Proposition \ref{Prop:ErrorExpansion}, let $\alpha\geq 1$. Assume $\{\gamma_k\}$ satisfies $\ensuremath {(\mathcal{H}_{\gamma})}$, and that $\Gamma_M^{[\alpha]}\rightarrow \infty$, for $\Gamma_M^{[\alpha]}$ defined as in \eqref{Eq:DefGammaAlpha}. Assume as well that the solution of the centered Poisson equation $\phi_\psi$ is in $C_{b,p}^{2,r}$ for $r> 3$. Let \(\bar{\Psi} := \int \psi(x,z)\mu^x(dz) \), then \begin{equation*} \EX{ \left| \frac{1}{\Gamma_M^{[\alpha]}}\sum_{k=1}^{M} \gamma^\alpha_k \left(\psi(x,\DecrAvg{Y}^{x,q}_{k-1}) -\bar{\Psi}(x)\right) \right|^2} \leq K \frac{ 1+ \Gamma^{[2\alpha-1]}_M +\Gamma^{[2\alpha]}_M + (\Gamma^{[\alpha+1]}_M)^2}{(\Gamma^{[\alpha]}_M)^2} \end{equation*} \end{MyProposition} \begin{proof} We recall first some martingale inequalities. Let $\{a_k\}$ be any sequence of random tensors. By Cauchy-Schwarz inequality we have that \begin{equation} \label{Eq:EstimateSquareGeneral} \EX{\left|\sum_{k=1}^{M}\gamma_k^p a_{k}\right|^2} \leq \EX{\Gamma^{[p]}_M \sum_{k=1}^{M}\gamma_k^p |a_{k}|^2} = \Gamma^{[p]}_M \sum_{k=1}^{M}\gamma_k^p \EX{|a_{k}|^2}. \end{equation} Let $\{b_k\}$ be also a sequence of tensors. If $s_0 < s_1 < \ldots < s_k < \ldots$, the $\{a_k\}, \{b_k\}$ are $\DecrAvg{\ensuremath \mathcal{F}}_{s_k}^q$ adapted and for all $k$, $\cEX{a_k}{\DecrAvg{\ensuremath \mathcal{F}}^q_{s_k}}=\cEX{b_k}{\DecrAvg{\ensuremath \mathcal{F}}^q_{s_k}}=0$, we have by martingale properties that \begin{equation} \EX{\left\langle \sum_{k=1}^{M}\gamma_k^p a_{k}, \sum_{k=1}^{M}\gamma_k^p b_{k}\right\rangle} = \sum_{k=1}^{M}\gamma_k^{2p}\EX{ \langle a_{k}, b_{k} \rangle }; \label{Eq:EstimateProductCentered} \end{equation} and in particular \begin{equation} \EX{\left|\sum_{k=1}^{M}\gamma_k^p a_{k}\right|^2} = \sum_{k=1}^{M}\gamma_k^{2p}\EX{ | a_{k} |^2}. \label{Eq:EstimateSquareCentered} \end{equation} Now, take the error expansion in Proposition \ref{Prop:ErrorExpansion} with $r=3$ and let $\eta_k = \gamma_k^\alpha$. By Abel's transformation, using convexity, the estimate \eqref{Eq:EstimateSquareGeneral}, the regularity properties of $\phi_\psi$ and Proposition \ref{Prop:ApproxFiniteMoments time}, we get \begin{equation} \label{Eq:ControlA0} \begin{split} & \EX{ |A_{\psi,M}^{0}(x,q) )|^2} = \EX{ \left| \sum_{k=1}^{M} \gamma_k^{\alpha-1} \left[ \phi_\psi^x(\DecrAvg{Y}^{x,q}_{k})-\phi_\psi^x(\DecrAvg{Y}^{x,q}_{k-1})\right]\right|^2}\\ & = \EX{ \left| \gamma_M^{\alpha-1}\phi_\psi^x(\DecrAvg{Y}^{x,q}_{M}) - \gamma_0^{\alpha-1}\phi_\psi^x(\DecrAvg{Y}^{x,q}_{0}) + \sum_{k=1}^{M-1} \left[(\gamma_k^{\alpha-1}-\gamma_{k+1}^{\alpha-1}) \phi_\psi^x(\DecrAvg{Y}^{x,q}_{k})\right]\right|^2}\\ & \quad \leq 3 \EX{ \left|\gamma_M^{\alpha-1}\phi_\psi^x(\DecrAvg{Y}^{x,q}_{M}) \right|^2} + 3 \EX{ \left|\gamma_0^{\alpha-1}\phi_\psi^x(\DecrAvg{Y}^{x,q}_{0}) \right|^2}\\ &\quad\quad + 3 \EX{ \left| \sum_{k=1}^{M-1} \left[(\gamma_k^{\alpha-1}-\gamma_{k+1}^{\alpha-1}) \phi_\psi^x(\DecrAvg{Y}^{x,q}_{k})\right]\right|^2}\\ & \leq K \left[(\gamma_M^{\alpha-1})^2 +1 + \left(\sum_{k=1}^{M-1}(\gamma_k^{\alpha-1}-\gamma_{k+1}^{\alpha-1}) \right)^2 \right] \leq K \end{split} \end{equation} Moreover using the fact that the terms are centered from Remark \ref{Rmk:ZeroExpectation}, equation \eqref{Eq:EstimateSquareCentered} and the finite moments of the Brownian increments imply \begin{gather} \EX{ |N_{{\psi},M} (x,q )|^2} = \sum_{k=1}^M \gamma_k^{2\alpha-1}\EX{ |\langle \sigma^* D_y \phi_{\psi} (x,\DecrAvg{Y}^{x,q}_{k-1}), U_k^q\rangle |^2} \leq K \Gamma_M^{2\alpha-1}\label{Eq:ControlN} \\ \begin{split} \EX{ |A_{{\psi},M}^{2}(x,q)|^2} &\leq \frac{1}{4} \sum_{k=1}^M \gamma_k^{2\alpha} \EX{ |D^2_y\phi_\psi(x,\DecrAvg{Y}^{x,q}_{k-1})\cdot(\sigma(x,\DecrAvg{Y}^{x,q}_{k-1})U^q_k)^{\otimes 2}) |^2} \\ &\leq K\Gamma^{[2\alpha]}_M \end{split} \label{Eq:ControlA2} \end{gather} More generally, estimate \eqref{Eq:EstimateSquareCentered} leads to \begin{equation} \label{Eq:ControlA3} \EX{ |A_{{\psi},M}^{3}(x,q)|^2} = \sum_{k=1}^M \gamma_k^{2\alpha+1} \EX{|v_{\psi}^{3,r}(x,\DecrAvg{Y}^{x,q}_{k-1},U_k^q)|^2} \leq K \Gamma_M^{[2\alpha+1]}, \end{equation} while by virtue of \eqref{Eq:EstimateSquareGeneral}, we find as estimate \begin{equation} \label{Eq:ControlZ3} \EX{ |Z_{{\psi},M}^{3}(x,q)|^2} \leq K \EX{\left|\sum_{k=1}^M \gamma_k^{\alpha+1} (1+|\DecrAvg{Y}^{x,q}_{k-1}|^{4})(1+|U_k^q|)^4\right|^2} \leq K (\Gamma_M^{[\alpha+1]})^2 \end{equation} On the other hand, from $\ensuremath {(\mathcal{H}_{\gamma})}$ and given that $\Gamma_M^{[\alpha]}\rightarrow \infty$, we have for $M$ large enough that, if $i>j$, \[\frac{\Gamma_M^{[i]}}{\Gamma^{[\alpha]}_M}\leq \frac{\Gamma_M^{[j]}}{\Gamma^{[\alpha]}_M}.\] The claim follows from Proposition \ref{Prop:ErrorExpansion} and \eqref{Eq:ControlA0},\eqref{Eq:ControlN},\eqref{Eq:ControlA2},\eqref{Eq:ControlA3},\eqref{Eq:ControlZ3}. \end{proof} \subsection{Ergodic average and Poisson equation} Being basic to our analysis, we introduce in this section some known properties of the exact averages and the effective diffusion. These results are studied in \cite{pardoux_poisson_2001} and \cite{pardoux_poisson_2003}. Let us start by stating a growth control result proved in \cite{veretennikov_polynomial_1997}. \begin{MyProposition}\label{Prop:FiniteMomentsTime} Let $\alpha>0$ and let $Y_t^{x}$ be the solution of \eqref{Eq:ErgodicDiffusion} with deterministic initial condition $y_0$ and coefficients satisfying $\ensuremath {(\mathcal{H}_{f.s.})}$. Then there exists a constant $K$ given only by $\alpha$, $\lambda_-$, $\lambda_+$ such that for all $t\geq 0$ and $x\in\ensuremath \mathds{R}^{d_x}$ \[\EX{ |Y_t^{x}|^\alpha} < K (1+|y_0|^{\alpha+2}) .\] \end{MyProposition} This proposition has a natural corollary \begin{MyCorollary}\label{Cor:FiniteMomentsIM} Under the same hypothesis of the theorem, for any $\alpha>0$ and all $x\in\ensuremath \mathds{R}^{d_x}$ \[\int |y|^\alpha \mu^x(dy) < K.\] \end{MyCorollary} \begin{MyLemma} \label{Lem:PropertiesAverage} Let $\psi(x,y)$ be a function satisfying the regularity and growth conditions in $\ensuremath {(\mathcal{H}_{s.s.})}$, and let $\Psi(x)=\int \psi(x,y) \mu^x(dy)$, then $\Psi(x)$ is $C^2_b$. \end{MyLemma} \begin{proof} The claim follows from adapting Theorems 3 and 5 in \cite{veretennikov_sobolev_2011} to the linear growth case: the needed equivalent results of convergence in total variation and control of expectations may be found in \cite{meyn_stability_1993}. \end{proof} As it was shown in Proposition \ref{Prop:ErrorExpansion}, the centered Poisson equation \eqref{Eq:CenteredPoisson} plays a special role in understanding the error expansion of the decreasing Euler algorithm. Proposition \ref{Prop:PropertiesPhi}, which is an adaptation of Theorem 1 in \cite{pardoux_poisson_2001} and \cite{veretennikov_sobolev_2011}, states some sufficient conditions for having the solution of such an equation when $f$ belongs to a certain family of functions. \begin{MyProposition} \label{Prop:PropertiesPhi} Consider a function $\psi(x,y)$ satisfying the regularity and growth conditions in $\ensuremath {(\mathcal{H}_{s.s.})} \ \ref{Enum:HipSlowScale_Lipschitz}), \ref{Enum:HipSlowScale_Growth})$; and such that \[\int \psi(x,y)\mu^x(dy) = 0, \forall x.\] Assume $\ensuremath {(\mathcal{H}_{f.s.})}$. Then, there exists a function $\phi_\psi(x,y)$, continuous in $y$ and belonging to the class $\bigcap_{p>1} W_{p,loc}^2$ in $y$, such that for every $x\in \ensuremath \mathds{R}^{d_x}$, \begin{enumerate}[i)] \label{Enum:SolvesPoisson} \item $\mathcal{L}^x_y \phi_\psi(x,y) = \psi(x,y)$, \label{Enum:IsCentered} \item $\int \phi_\psi(x,y) \mu^x(dy) = 0 $, \item $\phi_\psi \in C^{2,r^y}_{b,p}$. \end{enumerate} This function is the unique solution up to an additive constant of the Poisson equation on the class of continuous and $\bigcap_{p>1} W_{p,loc}^2$ functions in $y$ which are locally bounded and grow at most polynomially in $|y|$ as $|y|\rightarrow \infty$. Moreover, it has the representation \[\phi_\psi(x,y) = - \int_0^\infty \mathbb{E}_{x,y}( \psi(x,Y_t^x) ) dt \] \end{MyProposition} \section{Convergence results for the MsDS algorithm} \label{Sec:MsDSalgorithm} We focus now on the study of the MsDS algorithm. First, we show that the proposed approximated coefficients (by means of Decreasing Euler step and Cholesky procedures) satisfy a growth control and error control properties. As a consequence, we will conclude on some regularity property of the approximated diffusion \eqref{Eq:continuous_approximated} and show its strong convergence towards \eqref{Eq:EffectiveEquation}. Then, we will study the limit error distribution property. \subsection{Existence, uniqueness, continuity} From Hypothesis $\ensuremath {(\mathcal{H}_{s.s.})}$, $\ensuremath {(\mathcal{H}_{f.s.})}$, Proposition \ref{Prop:FiniteMomentsTime} and Proposition \ref{Prop:APriori}, it follows that there exists a unique solution to equation \ref{Eq:EffectiveEquation}, and that it has a continuous modification. We show the defined approximation has the same properties. Proposition \ref{Prop:ControlDecreasingStep} uses the results of section \ref{Sec:Preliminaries} to show that, under the standing hypothesis, the coefficients of the approximated diffusion have finite moments of any order, and that its error with respect to the exact coefficients decrease as a power of the number of steps $n$. \begin{MyProposition} \label{Prop:ControlDecreasingStep} Assume $\ensuremath {(\mathcal{H}_{s.s.})}$, $\ensuremath {(\mathcal{H}_{f.s.})}$ and $\ensuremath {(\mathcal{H}_{\gamma})}$. Let $\beta_0> 0$, and define $M(n)$ implicitly by $\Gamma_{M(n)} = C_0 n^{2\beta_0}$, where $C_0$ is some constant. \begin{enumerate}[i)] \item \label{Enum:ControlDecreasingStepPoisson} There exist $\phi_f$ and $\phi_h$ solutions of the centered Poisson equations \begin{itemize} \item $\mathcal{L}_y^x \phi_f(x,y) = f(x,y) - \int f(x,y')\mu^x(dy')$; and \item $\mathcal{L}_y^x \phi_h(x,y) = h(x,y) - \int h(x,y')\mu^x(dy').$ \end{itemize} \item \label{Enum:ControlDecreasingStepPibbmse} Let \begin{equation} \label{Eq:DefinitionLbar} \varsigma := \min_{l\geq 4, i=1,\ldots d} ( \bar{v}^{l,r^y}_{F^i} \neq 0 )\wedge \min_{l\geq 4, i,j=1,\ldots d} ( \bar{v}^{l,r^y}_{H^{i,j}} \neq 0 ) \wedge(r^y+1) \end{equation} (with the convention that $\min(\emptyset) = \infty)$ and $\bar{v}^{l,r}_{F^i}, \bar{v}^{l,r^y}_{H^{i,j}}$ defined as in \eqref{Eq:Definition_v_bar} applied to $F^1,\ldots,F^{d_x},H^{1,1},\ldots, H^{d_x,d_x} $. Assume the asymptotic expantion \begin{equation} \label{Eq:ConditionGammaMainDecreasingStep} \frac{\Gamma^{[\varsigma/2]}_M}{\Gamma_M} = C_1 n^{-\beta_1} + o(n^{-\beta_1}), \end{equation} for some $\beta_1>0$ and some constant $C_1$, holds. Let \begin{equation} \label{Eq:DefinitionBeta} \beta:=\beta_0\wedge\beta_1. \end{equation} Then $\DecrAvg{F}^n$ (and respectively $\DecrAvg{H}^n, \DecrAvg{G}^n:=\sqrt{\DecrAvg{H}^n}$) satisfies for any $\alpha\in \ensuremath \mathds{R}^+$ and $k=0,\ldots,n$ \begin{equation*} \begin{cases} \EX{|\DecrAvg{F}^n(x,t_k)| ^\alpha} \leq K \\ \EX{|\DecrAvg{F}^{n}(x,t_k) - F(x)|^2} \leq K n^{-2\beta}. \end{cases} \end{equation*} \end{enumerate} \end{MyProposition} \begin{MyRemark} We should understand $\varsigma$ as marking the first non-zero value in the error expansion of either $\DecrAvg{F}^n$ or $\DecrAvg{H}^n$. It depends exclusively on the coefficients of the effective and ergodic diffusion (in particular it does not depend on $n$). \end{MyRemark} \begin{MyRemark} Proposition \ref{Prop:ControlDecreasingStep} means that we have a rate of convergence in norm $L_2$ for the coefficient estimators of order $O(n^{-\beta})$. Since we \emph{choose} $\beta_0$ by taking $M(n)$ as needed, the actual limit to $\beta$ comes from $\beta_1$. But of course, increasing $\beta_0$ implies growing $M$ faster as a function of $n$, increasing the algorithm's cost. \end{MyRemark} \begin{proof}[Proof of Proposition \ref{Prop:ControlDecreasingStep}] Note first that $\ref{Enum:ControlDecreasingStepPoisson})$ follows from $\ensuremath {(\mathcal{H}_{s.s.})}$ and Proposition \ref{Prop:PropertiesPhi}. We prove $\ref{Enum:ControlDecreasingStepPibbmse})$. By Jensen inequality and Proposition \ref{Prop:ApproxFiniteMoments time} we have for every $\alpha\geq 1$ and $n$ big enough, \begin{multline*} \EX{|\DecrAvg{F}^n(x,q)|^\alpha} = \EX{\left|\frac{1}{\Gamma_M}\sum_{k=1}^M \gamma_k f(x,\DecrAvg{Y}^{x,q}_{k-1})\right|^\alpha} \\ \leq \EX{\frac{1}{\Gamma_M}\sum_{k=1}^M \gamma_k |f(x,\DecrAvg{Y}^{x,q}_{k-1})|^\alpha } \leq K, \end{multline*} and similarly for every $\alpha\geq 2$ \[\EX{|\DecrAvg{G}^n(x,q)|^\alpha} = \EX{|\DecrAvg{H}^n(x,q)|^{\alpha/2}} \leq K,\] since $|G|^2=|H|$. The result extends trivially to every $\alpha > 0$. It remains to prove the error control. We obtain an expansion of order $r^y$ in Proposition \ref{Prop:ErrorExpansion}. We can bound the first terms as we did in Proposition \ref{Prop:ControlDecreasingDifferentStep} by taking $\gamma_k=\eta_k$ for all $k=1,\ldots,M$ (i.e. taking $\alpha=1$ in the statement of Proposition \ref{Prop:ControlDecreasingDifferentStep}). More generally, from the definition of $\varsigma$ in \eqref{Eq:DefinitionLbar}, we have that for every $l<\varsigma$ or $l$ odd $\bar{v}^{l,r^y}_{F^{i}}(x,y)=0$ \eqref{Eq:EstimateSquareCentered} leads to \begin{equation} \label{Eq:ControlAi_centered} \EX{ |A_{{F^{i}},M}^{l}(x,q)|^2} = \sum_{k=1}^M \gamma_k^l \EX{|v_{F^i}^{l,r^y}(x,\DecrAvg{Y}^{x,q}_{k-1},U_k^q)|^2} \leq K \Gamma_M^{[l]}, \end{equation} while for even $l$ with $l\geq \varsigma$, by virtue of \eqref{Eq:EstimateSquareGeneral}, we find as estimate \begin{equation} \label{Eq:ControlAi_noncentered} \EX{ |A_{{F^{i}},M}^{l}(x,q)|^2} \leq \Gamma^{l/2}_M \sum_{k=1}^M \gamma_k^{l/2} \EX{|v_{F^i}^{l,r^y}(x,\DecrAvg{Y}^{x,q}_{k-1},U_k^q)|^2} \leq K (\Gamma_M^{[l/2]})^2, \end{equation} Likewise, \begin{equation} \label{Eq:ControlZi} \begin{split} \EX{ |Z_{{F^{i}},M}^{r^y}(x,q)|^2} & \leq K \EX{\left|\sum_{k=1}^M \gamma_k^{r^y+1/2} (1+|\DecrAvg{Y}^{x,q}_{k-1}|^{r^y+1})(1+|U_k^q|)^{r^y+1}\right|^2} \\\ & \leq K (\Gamma_M^{[r^y+1/2]})^2 \end{split} \end{equation} Note that estimates \eqref{Eq:ControlAi_centered} and \eqref{Eq:ControlAi_noncentered} are uniform in $x$. On the other hand, from $\ensuremath {(\mathcal{H}_{\gamma})}$, we have for $M$ big enough and $l\leq r^y$ that \[1\geq \frac{\Gamma_M^{[2]}}{\Gamma_M}\geq \frac{\Gamma_M^{[3]}}{\Gamma_M}\geq \ldots \frac{\Gamma_M^{[l]}}{\Gamma_M} ;\] Hence, from Proposition \ref{Prop:ErrorExpansion} and equations \eqref{Eq:ControlA0},\eqref{Eq:ControlN},\eqref{Eq:ControlA2},\eqref{Eq:ControlAi_centered},\eqref{Eq:ControlAi_noncentered}, \[\EX{|\DecrAvg{F}^{i;n}(x,q) -{F^{i}}(x,q)|^2}\leq \frac{ K (\Gamma^{[\varsigma/2]}_M)^2}{(\Gamma_M)^2} + \frac{K}{\Gamma_M} \leq K'n^{-2(\beta_0\wedge\beta_1)}.\] implying our claim for $F$, $\DecrAvg{F}^n$. Since $H$ satisfies the same properties than $F$, the claim follows for $H,\DecrAvg{H}^{n}$. As a final step, we prove the error control for $\DecrAvg{G}^n$. Let $\Delta H^n(x,q):=H(x)-\DecrAvg{H}^n(x,q)$ and \(E = \left\{|\Delta H^n(x,q)|\geq |2H^{-1}|^{-1}\right\}.\) Markov inequality gives us the control \begin{align*} \probX{E} &\leq 4 |H^{-1}(x)|^2 \EX{|\Delta H^n(x,q)|^2} \leq K n^{-2(\beta_0 \wedge \beta_1)} \end{align*} which, in conjunction with Theorem \ref{Theo:ControlCholesky}, deduces \begin{align*} &\EX{|G(x)-\DecrAvg{G}^n(x,q)|^2} \\ & \quad = \EX{|G(x)-\DecrAvg{G}^n(x,q)|^2\mathbf{1}_{E}} + \EX{|G(x)-\DecrAvg{G}^n(x,q)|^2\mathbf{1}_{E^\complement}}\\ & \quad \leq K^\prime n^{-2(\beta_0 \wedge \beta_1)} + \EX{|G(x)-\DecrAvg{G}^n(x,q)|^2\mathbf{1}_{E^\complement}}\\ & \quad \leq K^{\prime} n^{-2(\beta_0 \wedge \beta_1)} + K n^{-2(\beta_0 \wedge\beta_1)} = K^{\prime\prime} n^{-2(\beta_0 \wedge \beta_1)}. \end{align*} \end{proof} We can deduce from Proposition \ref{Prop:ControlDecreasingStep} and the assumed structure the following a priori estimates. \begin{MyCorollary} \label{Cor:ControlPowerAndDifference} Under the hypothesis and notation of Proposition \ref{Prop:ControlDecreasingStep}, for any $0\leq s\leq T$, \begin{equation} \label{Eq:BoundExpectedPower} \EX{|\DecrAvg{F}^n(\DecrAvg{X}^n_{\tbar[s]},\tbar[s])| ^\alpha} \leq K, \end{equation} and \begin{equation} \label{Eq:BoundMeanSquareError} \EX{|\DecrAvg{F}^n(\DecrAvg{X}_s^n,\tbar[s]) - F(\DecrAvg{X}_s^n)|^2} \leq K n^{-2\beta}. \end{equation} The same bounds hold with $\DecrAvg{F}^n, F$ replaced by $\DecrAvg{H}^n,H$ and $\DecrAvg{G}^n,G$. \end{MyCorollary} \begin{proof} Define \begin{equation} \label{Eq:DefinitionFttminusFilter} \bar{\ensuremath \mathcal{F}}_{t,t^-}:=\left(\ensuremath \mathcal{F}_t \vee \bigvee\limits_{q\in \ensuremath \mathds{Q}, q < t} \tilde{\ensuremath \mathcal{F}}^q_\infty \right) \end{equation} by construction, $\ApproxVar{X}_{\tbar[s]}$ is $\bar{\ensuremath \mathcal{F}}_{\tbar[s],\tbar[s]^-}$ measurable and since $\DecrAvg{F}^n(x,\tbar[s]) \protect\mathpalette{\protect\independenT}{\perp} \bar{\ensuremath \mathcal{F}}_{\tbar[s],\tbar[s]^-}$ for any deterministic $x$, we get from Proposition \ref{Prop:ControlDecreasingStep} \[\EX{|\DecrAvg{F}^n(\ApproxVar{X}_{\tbar[s]},\tbar[s])| ^\alpha} =\EX{\cEX{|\DecrAvg{F}^n(\ApproxVar{X}_{\tbar[s]},\tbar[s])| ^\alpha}{\bar{\ensuremath \mathcal{F}}_{\tbar[s],\tbar[s]^-}}} \leq \EX{K} = K.\] A similar argument leads to \eqref{Eq:BoundMeanSquareError}, and to the claims for $\DecrAvg{H}^n,H$ and $\DecrAvg{G}^n,G$. \end{proof} Corollary \ref{Cor:ControlPowerAndDifference} should be understood as an a priori control on the approximated process. From this control, we can deduce, using Proposition \ref{Prop:APriori} as in the case of the effective equation, the existence and strong uniqueness of the solution of the approximated diffusion \eqref{Eq:continuous_approximated}. In addition, Proposition \ref{Prop:EstimatePowerApprox} states that the approximation \eqref{Eq:continuous_approximated} has a continuous modification. The result follows from Proposition \ref{Prop:FiniteMomentsTime}, the estimates in Corollary \ref{Cor:ControlPowerAndDifference} and Kolmogorov's criterion. \begin{MyProposition} \label{Prop:EstimatePowerApprox} Under the hypothesis and notation of Proposition \ref{Prop:ControlDecreasingStep}, for every $\alpha\geq 2$ \[\EX{|\ApproxVar{X}^{n}_t - \ApproxVar{X}^{n}_s|^\alpha} \leq K_{\alpha,T} (t-s)^{\alpha/2} ((t-s)^{\alpha/2} +1 ).\] Moreover, the solution of \eqref{Eq:continuous_approximated} has a continuous modification. \end{MyProposition} \subsection{Strong convergence} \label{SubSec:Almostsurepathwise} In what follows, we choose $\ApproxVar{X}$ to be continuous in time. We can proceed to show the mean square convergence of $\ApproxVar{X}^{n}$ towards $X$. \begin{MyTheorem} \label{Theo:GeneralConvergence} Under $\ensuremath {(\mathcal{H}_{s.s.})}$, $\ensuremath {(\mathcal{H}_{f.s.})}$ and $\ensuremath {(\mathcal{H}_{\gamma})}$, let $X$ be defined by \eqref{Eq:EffectiveEquation} and $\ApproxVar{X}^{n}$ by \eqref{Eq:continuous_approximated}. Let $\beta$ be defined as in \eqref{Eq:DefinitionBeta}. Then, \begin{itemize} \item If $g\equiv 0$ (ODE with random coefficients), then \\ $\EX{ \sup\limits_{0\leq t \leq T} |X_t-\ApproxVar{X}^{n}_t|^2} \leq K n^{-2(1\wedge \beta)}.$ \item Under the full SDE case, $\EX{ \sup\limits_{0\leq t \leq T} |X_t-\ApproxVar{X}^{n}_t|^2} \leq K n^{-(1\wedge 2\beta)}.$ \end{itemize} \end{MyTheorem} \begin{proof} We treat the full SDE case. By definition \begin{equation*} X_{t} - \ApproxVar{X}^{n}_{t} = \int_0^{t} \left[F (X_{s}) - \DecrAvg{F}^{n}(\ApproxVar{X}^{n}_{\tbar[s]},\tbar[s]) \right]ds + \int_0^{t} \left[ G (X_{s}) -\DecrAvg{G}^{n}(\ApproxVar{X}^{n}_{\tbar[s]},\tbar[s])\right]dW_s. \end{equation*} Our plan is to use Proposition \ref{Prop:APriori} $\ref{Enum:AprioriControl})$. By convexity \begin{align*} & \left|F (X_{s}) - \DecrAvg{F}^{n}\left(\ApproxVar{X}^{n}_{\tbar[s]},\tbar[s]\right) \right|^2 \\ & \quad\quad\quad \leq 3 \left|F (X_{s}) - F\left(\ApproxVar{X}^{n}_s\right) \right|^2 + 3\left|F\left(\ApproxVar{X}^{n}_s\right) - F\left(\ApproxVar{X}^{n}_{\tbar[s]}\right) \right|^2\\ & \quad\quad\quad + 3\left|F\left(\ApproxVar{X}^{n}_{\tbar[s]},\tbar[s]\right) - \DecrAvg{F}^{n}\left(\ApproxVar{X}^{n}_{\tbar[s]},\tbar[s]\right) \right|^2. \end{align*} By Lipschitz assumption in $\ensuremath {(\mathcal{H}_{s.s.})}$, \begin{align} \EX{\left|F (X_{s}) - F\left(\ApproxVar{X}^{n}_s\right) \right|^2} & \leq K \EX{\left|X_{s} - \ApproxVar{X}^{n}_s \right|^2} \notag \\ \EX{\left|F\left(\ApproxVar{X}^{n}_s\right) - F\left(\ApproxVar{X}^{n}_{\tbar[s]}\right) \right|^2} & \leq K \EX{\left|\ApproxVar{X}^{n}_s - \ApproxVar{X}^{n}_{\tbar[s]}\right|^2} \leq K n^{-1}, \label{Eq:EulerError} \end{align} the last inequality being possible for $n$ large enough thanks to Proposition \ref{Prop:EstimatePowerApprox}. Also, by Corollary \ref{Cor:ControlPowerAndDifference} , we get \[ \EX{ \left| F\left(\ApproxVar{X}^{n}_{\tbar[s]},\tbar[s]\right) - \DecrAvg{F}^{n}\left(\ApproxVar{X}^{n}_{\tbar[s]},\tbar[s]\right) \right|^2 } \leq K n^{-2\beta}.\] Therefore \begin{equation} \label{Eq:BoundF} \EX{ \left|F (X_{s}) - \DecrAvg{F}^{n}\left(\ApproxVar{X}^{n}_{\tbar[s]},\tbar[s]\right) \right|^2} \leq K \left(n^{-(1\wedge 2\beta)} + \EX{|X_s-\ApproxVar{X}^{n}_s|^2}\right), \end{equation} Since we may obtain similar bounds for the terms with $G$, we also have \begin{equation} \label{Eq:BoundG} \EX{ \left|G (X_{s}) - \DecrAvg{G}^{n}\left(\ApproxVar{X}^{n}_{\tbar[s]},\tbar[s]\right) \right|^2} \leq K \left(n^{-(1\wedge 2\beta)} + \EX{|X_s-\ApproxVar{X}^{n}_s|^2}\right). \end{equation} Now, Proposition \ref{Prop:APriori} $\ref{Enum:AprioriControl})$ shows \begin{align*} \EX{|X_{t}-\ApproxVar{X}^{n}_{t}|^2} & \leq K \int_0^{T} \left(n^{-(1\wedge 2\beta)} + \EX{|X_s-\ApproxVar{X}^{n}_s|^2}\right)ds, \end{align*} Therefore, by Gronwall's lemma, \[\sup\limits_{0\leq t \leq T} \EX{|X_t-\ApproxVar{X}^{n}_t|^2} \leq K n^{-(1\wedge 2\beta)}.\] Replacing on \eqref{Eq:BoundF} and \eqref{Eq:BoundG} we get \begin{multline*} \sup\limits_{0\leq t \leq T} \left( \EX{ \left|F (X_{s}) - \DecrAvg{F}^{n}\left(\ApproxVar{X}^{n}_{\tbar[s]},\tbar[s]\right) \right|^2} + \EX{ \left|G (X_{s}) - \DecrAvg{G}^{n}\left(\ApproxVar{X}^{n}_{\tbar[s]},\tbar[s]\right) \right|^2}\right)\\ \leq K n^{-(1\wedge 2\beta)}. \end{multline*} So that by Proposition \ref{Prop:APriori} $\ref{Enum:AprioriControl})$, \begin{equation*} \EX{ \sup\limits_{0\leq t \leq T} |X_t-\ApproxVar{X}^{n}_t|^2} \leq K n^{-(1\wedge 2\beta)}. \end{equation*} Note that the case $g\equiv 0$ is proven in the same way, but the Euler error \eqref{Eq:EulerError} is bounded by $n^{-2}$ and $G\equiv 0$. This implies the stated result. \end{proof} \subsection{Limit distribution} \label{SubSec:LimitDistribution} In this section we show under slightly stronger regularity assumptions on the coefficients of the diffusion, that we have convergence in the weak (uniform topology) sense towards a limit distribution given as the solution of a particular SDE. Our plan to prove the limit distribution result is to look at the rescaled error and its associated stochastic differential equation. We prove the joint weak convergence of the terms appearing in that SDE and use the fact that under certain hypothesis the joint convergence of the terms suffices to deduce the weak convergence of the solution of the equation. The reader may find most of the needed material on weak convergence of stochastic integrals and stochastic SDEs in \cite{jakubowski_convergence_1989}, \cite{kurtz_weak_1991} and \cite{kurtz_weak_1996}. \begin{MyDefinition} Let $X^{n}$ be a sequence of $\ensuremath \mathds{R}^d$-valued semimartingales and let $A^{n}(\delta)$ be the predictable process with finite variation null at zero and $M^{n}(\delta)$ the local martingale null at zero appearing in the representation of $X^{n}$ as \[X^{n}_t = X^{n}_{0} + A^{n}_t(\delta) + M^{n}_t(\delta) + \sum_{s\leq t}\Delta X_s^{n}\one{|\Delta X_s^n|>\delta}.\] We say that the sequence $X^{n}$ satisfies property \eqref{Property:Star} if for some $\delta > 0$ \[ \langle M^{n}(\delta),M^{n}(\delta) \rangle_T + \int_0^T |dA^{n}(\delta)_s|+ \sum_{s\leq T} |\Delta X_s^{n}|\one{|\Delta X_s^{n}|>\delta} \tag{*} \label{Property:Star}\] is tight. (The notation $\int_0^T |dA|$ denotes the total variation of $A$ on $[0,T]$) \end{MyDefinition} The importance of property \eqref{Property:Star} is shown by the following theorem (see \cite{jakubowski_convergence_1989}, \cite{jacod_asymptotic_1998} and \cite{kurtz_weak_1996}). \begin{MyTheorem} \label{Theo:StarConvergenceIntegral} Let $X^{n}$ be a sequence of $\ensuremath \mathds{R}^d-$ valued semimartingales relative to the filtration $\mathcal{F}_t$. Suppose that $X^{n}$ weakly converges in the Skorokhod topology $D_{\ensuremath \mathds{R}^{d_x}}$. Then \eqref{Property:Star} is necessary and sufficient for goodness: for any sequence $H^n$ of $(\ensuremath \mathcal{F}_t)-$ adapted c\`{a}dl\`{a}g processes such that $(H^n,X^n)\Rightarrow (H,X)$ in the Skorokhod topology $D_{M^{d_x\times d_x}\times\ensuremath \mathds{R}^{d_x}}$, then $X$ is a semimartingale w.r.t the filtration generated by $(H,X)$ and $(H^n,X^n,\int H^n dX^n) \Rightarrow (H,X,\int H dX)$ in the Skorokhod topology $D_{M^{d_x\times d_x}\times\ensuremath \mathds{R}^{d_x}\times\ensuremath \mathds{R}^{d_x}}$. \end{MyTheorem} Goodness gives us a direct way to show the convergence of sequences of stochastic integrals, and will play a key role for the convergence of sequences of SDEs. Before proceeding to the main propositions of this section, we cite another useful result concerning weak convergence of sequences of solutions of SDEs, allowing to compare the limit of two sequences with converging coefficients. \begin{MyTheorem}[Theorem 2.5 (b) \cite{jacod_asymptotic_1998}] \label{Thm:ConvergenceChangeCoefficients} Consider a sequence of linear SDEs \begin{equation} \label{Eq:LinearSDE} \vartheta_t^n = P_t^n +\int_0^t \vartheta^n_{s-} Q_t^n dJ_t \end{equation} where the $P_t^n $ are stochastic processes in $\ensuremath \mathds{R}^d$, $Q_t^n $ are stochastic processes in $\ensuremath \mathds{R}^{d\times d'}$ and $J_t$ is a semimartingale in $\ensuremath \mathds{R}^{d'}$, and all processes are in same the filtered probability space. Suppose that we have another sequence of equations like \eqref{Eq:LinearSDE} with solution $\vartheta^{\prime n}$ and coefficients $P^{\prime n}$ and $Q^{\prime n}$. If the sequences $ \sup_{0\leq s \leq T } \|P^{n}_s\|$ and $\sup_{0\leq s \leq T } \|Q^{n}_s\| $ are tight and if \begin{align*} &\sup_{0\leq s \leq T } \|P^{n}_s - P^{\prime n}_s\| \tendsprob 0 &\sup_{0\leq s \leq T } \|Q^{n}_s - Q^{\prime n}_s\| \tendsprob 0 \end{align*} then, \[\sup_{0\leq s \leq T } \|\vartheta^{n}_s - \vartheta^{\prime n}_s\| \tendsprob 0\] \end{MyTheorem} Proposition \ref{Prop:ConvergenceOfTuple} shows the weak convergence of some tuples appearing in the rescaled error SDE. \begin{MyProposition} \label{Prop:ConvergenceOfTuple} Let $\mathcal{I}$ be a set of indices, and consider a family of independent standard Gaussian variables $ \{\nu^{i;n}_{t_k}\}_{n\in\ensuremath \mathds{N}^*; 0\leq k \leq n; i\in \mathcal{I}}$ where for any $n,i$ we have $\nu^{i;n}_{t_k}$ is $\bar{\ensuremath \mathcal{F}}_{t_k}$ measurable. Consider the sequence of random processes $A^{0;n}$ ( dimension 1), $A^{1;n},B^{0;n}$ (dimension $d_x$), $B^{2;n}$ (dimension $d_x\times d_x$), $B^{1;n}$ (dimension $|\mathcal{I}|$) and $B^{3;n}$ (dimension $|\mathcal{I}|\times d_x$) defined component-wise by \begin{align} \label{Eq:DefHc} B^{0;j;n}_t &:= \int_0^t (s-\tbar[s])dW^j_s; & A^{0;n}_t &:= 2\int_0^t (s-\tbar[s])ds ; \\ \label{Eq:DefJc} B_t^{2;l,j;n} &:= \int_0^t \sqrt{2} (W^l_s - W^l_{\tbar[s]}) d W^j_s; & A^{1;j;n}_t &:= \int_0^t (W^j_s - W^j_{\tbar[s]}) ds; \\ \label{Eq:DefKc} B_t^{3;i,j;n} &:= \int_0^t \nu^{i;n}_{\tbar[s]} dW^j_s; & B_t^{1;i;n} &:= \int_0^t \nu^{i;n}_{\tbar[s]} ds ; \end{align} Then, we have the following limit results \begin{equation} (X,\ApproxVar{X}^{n},W,n A^{0;n}, \sqrt{n} B^{1;n} ) \Rightarrow (X,X,W,A^{0},B^{1}) \label{Eq:TupleConv1} \end{equation} \begin{multline} (X,\ApproxVar{X}^{n}, W,n^{\frac{1}{2}}A^{0;n},n^{\frac{1}{2}}B^{0;n}, n^{\frac{1}{2}}A^{1;n}, n^{\frac{1}{2}}B^{2;n}_s,B^{1;n},B^{3;n}) \\ \Rightarrow (X,X,W,0,0,0,B^{2},0,B^{3}) \label{Eq:TupleConv3} \end{multline} where $A^{0}_t=t$; $B^{0}$, $B^{1}$, $B^{2}$ and $B^{3}$ are standard Brownian motions defined on an extension of the space $W$, with dimensions $d_x$, $d_x^2$, $|\mathcal{I}|\times d_x$ and $|\mathcal{I}|$ respectively. Moreover, we have $\{ B^{0}$, $B^{2}$, $B^{3}$, $W\}$ are independent; $\{ B^{0}$, $B^{2}$, $B^{1}$, $W\}$ are independent, and $B^{1;n}$, $\sqrt{n}B^{2;n}$ and $B^{3;n}$ are ``good'' in the sense of Theorem \ref{Theo:StarConvergenceIntegral}. \end{MyProposition} The proof of Proposition \ref{Prop:ConvergenceOfTuple} will be given in section \ref{SubSec:WeakConvTuples}. \begin{MyProposition} \label{Prop:MainDecreasingStepPqCLT} Under the assumptions and notation of Proposition \ref{Prop:ControlDecreasingStep}, assume that $r^y>\varsigma+3$ in $\ensuremath {(\mathcal{H}_{s.s.})}$, and that there is $\beta_2 \geq 0 $ such that the asymptotic expansion \begin{equation} \label{Eq:ConditionGammaMainDecreasingStepLimitDist} \frac{\Gamma^{[\varsigma/2+1]}_M}{\Gamma_M^{[\varsigma/2]}} = C_2 n^{-\beta_2} + o(n^{-\beta_2}), \end{equation} where $\varsigma$ is defined in \eqref{Eq:DefinitionLbar}, holds. Let \begin{equation} \label{Eq:DefinitionRho} \rho = \one{\beta_0>\beta_1} (\beta_2 \wedge (\beta_0-\beta_1)) + \one{\beta_0<\beta_1} (\beta_0 \wedge (\beta_1-\beta_0)). \end{equation} \begin{enumerate}[i)] \item \label{Enum:LimitDistPart1}Let $\Phi_F$ be the $d_x\times d_x$ matrix defined component-wise as \begin{align*} \Phi_F^{i,j}(x) &:= C_0^{-1}\int \langle \sigma^* D_y \phi_{F^i}(x,y), \sigma^* D_y \phi_{F^j}(x,y) \rangle \mu^x(dy), \end{align*} where $\phi_{F^i}$ is the solutions of the Poisson equation \eqref{Eq:CenteredPoisson} with source $F^i$. Let \begin{align*} \varphi_F(x) &:= \one{\beta_1 \geq \beta_0} \sqrt{\Phi_F(x)}; & R_{F}^i(x) &:= \one{\beta_0\geq \beta_1} {C_1}\int \bar{v}_{F^i}^{\varsigma,r^y}(x,y) \mu^x(dy) \end{align*} with the square root meaning the Cholesky root. Then, there exist a family of independent standard Gaussian variables $ \{\nu^{i;n}_{k}\}_{n\in\ensuremath \mathds{N}^*; 0\leq k \leq n; 1\leq i \leq d_x}$, such that each $\nu^{i;n}_{k}$ is $\bar{\ensuremath \mathcal{F}}_{t_k}$ measurable and \begin{equation*} \EX{\left| n^\beta \left( F^i(x)-\DecrAvg{F}^{i;n}(x,t_k) \right) - \sum_{j=1}^{d_x} \varphi_F^{i,j}(x) \nu^{j;n}_{k} - R_{F}^i(x) \right|^2} = O(n^{-2\rho}),\\ \end{equation*} for all $x\in \ensuremath \mathds{R}^{d_x}$. \item Under the full SDE case, define in a similar way a $d_x^2$ dimensional random function $R_H$ and a $d_x^2\times d_x^2$ dimensional random function $\Phi_H$, with \begin{align*} \Phi_H^{i,j,i',j'}(x) &:= C_0^{-1}\int \langle \sigma^* D_y \phi_{H^{i,j}}(x,y), \sigma^* D_y \phi_{H^{i',j'}}(x,y) \rangle \mu^x(dy), \end{align*} \begin{align*} \varphi_H(x) &:= \one{\beta_1\geq \beta_0} \sqrt{\Phi_H(x)}; & R_{H}^{i,j}(x) &:= \one{\beta_0 \geq \beta_1} {C_1}\int \bar{v}_{H^{i,j}}^{\varsigma,r^y}(x,y) \mu^x(dy) \end{align*} Then, there exist a family of independent standard Gaussian variables $ \{\nu^{i,j;n}_{k}\}_{n\in\ensuremath \mathds{N}^*; 0\leq k \leq n; 0\leq i,j \leq d_x}$, such that each $\nu^{i,j;n}_{k}$ is $\bar{\ensuremath \mathcal{F}}_{t_k}$ measurable and \begin{multline*} \EX{\left| n^\beta \left( H^{i,j}(x)-\DecrAvg{H}^{i,j;n}(x,t_k) \right) - \sum_{i',j'=1}^{d_x} \varphi_H^{i,j,i',j'}(x) \nu^{i',j';n}_{k} - R_{H}^{i,j}(x) \right|^2}\\ = O(n^{-2\rho}), \end{multline*} for all $x\in \ensuremath \mathds{R}^{d_x}$. Moreover, letting $R_G$, $\varphi_G$ be defined component-wise for $0\leq i',j'\leq d_x $ as \begin{align*} R_{G}^{i,i} &= \frac{ R_{H}^{i,i} - 2 \sum\limits_{k=1}^{i-1} R_{G}^{i,k} G^{i,k}}{2G^{i,i}} , & \varphi_{G}^{i,i,i',j'} & = \frac{ \varphi_{H}^{i,i,i',j'} - 2 \sum\limits_{j=1}^{i-1} \varphi_{G}^{i,j,i',j'}G^{i,j}}{2 G^{i,i}} ; \end{align*} and for $i>j$ \begin{gather*} R_{G}^{i,j} = \frac{R_{H}^{i,j} -R_{G}^{j,j} G^{i,j} -\sum\limits_{l=1}^{j-1} (R_{G}^{j,l} G^{i,l}+R_{G}^{i,l}G^{j,l})}{G^{j,j}} \\ \begin{split} \varphi_{G}^{i,j,i',j'} = \frac{ \varphi_{H}^{i,j,i',j'} - \varphi_{G}^{j,j,i',j'} G^{i,j}-\sum\limits_{l=1}^{j-1} \left[\varphi_{G}^{j,l,i',j'} G^{i,l}+\varphi_{G}^{i,l,i',j'} G^{j,l}\right] }{G^{j,j}}. \end{split} \end{gather*} \end{enumerate} Then, \begin{multline*} \EX{\left| n^\beta \left( G^{i,j}(x)-\DecrAvg{G}^{i,j;n}(x,t_k) \right) - \sum_{i',j' =1}^{d_x} \varphi_G^{i,j,i',j'}(x) \nu^{i',j';n}_{k} - R_G^{i,j}(x) \right|^2} \\= O(n^{-2\rho}). \end{multline*} \end{MyProposition} \begin{proof} \begin{enumerate}[i)] \item We prove the first claim. We use the expansion of Proposition \ref{Prop:ErrorExpansion} up to order $\varsigma$ as in Proposition \ref{Prop:ControlDecreasingStep}, and estimates \eqref{Eq:ControlA0}, \eqref{Eq:ControlN}, \eqref{Eq:ControlA2}, \eqref{Eq:ControlAi_centered}, \eqref{Eq:ControlAi_noncentered}, \eqref{Eq:ControlZi} ; to get for any $x$ that \begin{equation} \label{Eq:MainErrorTerms} \begin{split} \EX{\left| ({F^i}(x)-\DecrAvg{{F}}^{i;n}(x,q)) - \frac{1}{{\Gamma_M}} \left(N_{{F^i},M}(x,q)+A_{{F^i},M}^{(\varsigma)}(x,q)\right) \right|^2} \\ \quad = O\left( (\Gamma_M)^{-2} \left[1 +\left(\Gamma_M^{[\varsigma/2+1]}\right)^2\right]\right) \end{split} \end{equation} Let us examine separately three cases depending on the relation between $\beta_0$ and $\beta_1$: \begin{itemize} \item If $\beta_0>\beta_1$: In this case $\beta=\beta_1$, and by definition of $\beta_1$ it follows that \begin{align} &\EX{\left| n^\beta \left( F^i(x)-\DecrAvg{F}^{i;n}(x,q) \right) - R_{F}^i(x) \right|^2} \label{Eq:ExpansionLimitBigBeta0}\\ & \ \leq K \EX{\left| (\Gamma^{[\varsigma/2]})^{-1}\Gamma_M \left( F^i(x)-\DecrAvg{F}^{i;n}(x,q) \right) - R_{F}^i(x) \right|^2} \notag\\ & \ \leq K' \EX{|\Gamma^{[\varsigma/2]}_M|^{-2} \left|\Gamma_M(F^i(x)-\DecrAvg{F}^{i;n}(x,q)) - N_{F^i,M}(x,q)- A_{F^i,M}^{\varsigma}(x,q) \right|^2} \notag \\ & \quad + K' \EX{ \left|( \Gamma^{|[\varsigma/2]}_M)^{-1} (A_{F^i,M}^{\varsigma}(x,q) - R_{F}^i(x))\right|^2 } \notag\\ & \quad + K' \EX{ \left|( \Gamma^{|[\varsigma/2]}_M)^{-1} N_{F^i,M}(x,q) \right|^2 } \notag \end{align} The first term in the right hand side of \eqref{Eq:ExpansionLimitBigBeta0} can be controlled by rescaling \eqref{Eq:MainErrorTerms} to get \begin{multline} \EX{|\Gamma^{[\varsigma/2]}_M|^{-2} \left|\Gamma_M(F^i(x)-\DecrAvg{F}^{i;n}(x,q)) - N_{F^i,M}(x,q)- A_{F^i,M}^{\varsigma}(x,q) \right|^2} \\ = O\left( (\Gamma_M^{[\varsigma/2]})^{-2} \left[ \left(\Gamma_M^{[\varsigma/2+1]}\right)^2 + 1 \right] \right). \label{Eq:ErrorPsiB0greaterB1} \end{multline} From \eqref{Eq:ControlN} we control the third term in the right hand side of \eqref{Eq:ExpansionLimitBigBeta0} \begin{equation} \label{Eq:OrderNPsi} \EX{\left|(\Gamma_M^{[\varsigma/2]})^{-1} N_{F^i,M}(x,q)\right|^2} = O\left( (\Gamma_M^{[\varsigma/2]})^{-2} \Gamma_M\right). \end{equation} To control the second term of \eqref{Eq:ExpansionLimitBigBeta0}, let us define \begin{equation} \label{Eq:DefinitionAbar} \bar{A}_{F^i,M}^{\varsigma}(x,q):= \sum_{k=1}^{M} \gamma_k^{\varsigma/2} \bar{v}^{\varsigma,r^y}_{F^i}(x,\bar{Y}^{x,q}_{k-1}) \end{equation} for $\bar{v}^{\varsigma,r^y}_{F^i}$ defined in \eqref{Eq:Definition_v_bar}. We can compare $(\Gamma^{[\varsigma/2]}_M)^{-1} A^{\varsigma}_{F^i,M}$ and $(\Gamma^{[\varsigma/2]}_M)^{-1} \bar{A}^{\varsigma}_{F^i,M}$ in $L_2$ by \eqref{Eq:EstimateSquareCentered}. Indeed, thanks to controls \eqref{Eq:ControlAi_centered} and \eqref{Eq:ControlAi_noncentered}, and the fact that for some $K\in \ensuremath \mathds{R}^+$, $\Gamma_M^{[\varsigma]} \leq K \Gamma_M$, we have \begin{multline} \label{Eq:AminusAbar} \EX{\left| (\Gamma^{[\varsigma/2]}_M)^{-1} \left( A^{\varsigma}_{F^i,M}(x,q) - \bar{A}^{\varsigma}_{F^i,M}(x,q)\right) \right|^2} \\ \quad\quad = \EX{\left| (\Gamma^{[\varsigma/2]}_M)^{-1}\sum_{k=1}^{M}\gamma_k^{\varsigma/2}\left( v_{F^i}^{\varsigma,r^y}(x,\bar{Y}^{x,q}_{k-1},U_k^q) - \bar{v}^{\varsigma,r^y}_{F^i}(x,\bar{Y}^{x,q}_{k-1})\right) \right|^2}\\ = O\left((\Gamma^{[\varsigma/2]})^{-2} \Gamma_M\right). \end{multline} It remains to show that \begin{equation} \label{Eq:AbarminusR}\EX{\left|(\Gamma^{[\varsigma/2]}_M)^{-1}\bar{A}_{F^i,M}^{\varsigma}(x,q) + R_F^i(x) \right|^2}=O(n^{-2\rho}).\end{equation} Indeed, from the definition of $\beta_0$ and $\beta_1$, $\Gamma^{[\varsigma/2]}_M = O(n^{\beta_0-\beta_1})$ so that it diverges. Moreover, from the assumed regularity hypothesis, $\bar{v}^{\varsigma,r^y}_{F^i}(x,y)$ is $C_{p,b}^{2,r^y-\varsigma}$. Therefore, Proposition \ref{Prop:PropertiesPhi} guarantees the existence of a solution to the centered Poisson equation with source $\bar{A}^{\varsigma,r^y}_{F^i}(x,y)$ of the same regularity, and thus Proposition \ref{Prop:ControlDecreasingDifferentStep} shows that $\bar{A}^{\varsigma}_{F^i,M}(x,q)$ converges uniformly with respect to $x$ in $L^2$ to $-R_F^i(x)$ with rate $(\beta_0-\beta_1) \wedge \beta_2 \geq \rho$ since \begin{align*}&(\Gamma_M^{\varsigma/2})^{-2} \left(1+\Gamma_M^{[\varsigma]}+\Gamma_M^{[\varsigma-1]}+(\Gamma_M^{[\varsigma/2+1]})^2\right) \\ & \quad\quad \leq K (\Gamma_M^{\varsigma/2})^{-2}\left(\Gamma_M+(\Gamma_M^{[\varsigma/2+1]})^2\right)= O(n^{-2((\beta_0-\beta_1 )\wedge \beta_2)}). \end{align*} The claim follows from replacing \eqref{Eq:ErrorPsiB0greaterB1}, \eqref{Eq:OrderNPsi}, \eqref{Eq:AminusAbar} and \eqref{Eq:AbarminusR} in \eqref{Eq:ExpansionLimitBigBeta0}. \item If $\beta_1>\beta_0$: We follow a similar approach. We expand the rescaled error term to find \begin{align} &\EX{\left| n^\beta \left( F^i(x)-\DecrAvg{F}^{i;n}(x,q) \right) - R_{F}^i(x) \right|^2} \label{Eq:ExpansionLimitBigBeta1}\\ & \ \leq K' \EX{|\Gamma_M|^{-1/2} \left|\Gamma_M(F^i(x)-\DecrAvg{F}^{i;n}(x,q)) - N_{F^i,M}(x,q)- A_{F^i,M}^{\varsigma}(x,q) \right|^2} \notag \\ & \quad + K' \EX{ \left|( \Gamma_M)^{-1/2} A_{F^i,M}^{\varsigma}(x,q) \right|^2 } \notag\\ & \quad + K' \EX{ \left|( \Gamma_M)^{-1/2} (N_{F^i,M}(x,q) - \Phi_F)\right|^2 } \notag \end{align} By rescaling \eqref{Eq:MainErrorTerms} we get \begin{multline*} \EX{|\Gamma_M|^{-1/2} \left|\Gamma_M(F^i(x)-\DecrAvg{F}^{i;n}(x,q)) - N_{F^i,M}(x,q)- A_{F^i,M}^{\varsigma}(x,q) \right|^2} \\ = O\left( (\Gamma_M)^{-1} \left[ 1+\left(\Gamma_M^{[\varsigma/2+1]}\right)^2 \right]\right) \end{multline*} And from \eqref{Eq:ControlAi_noncentered} \[\EX{\left|(\Gamma_M)^{-1/2} A_{F^i,M}^{\varsigma}(x,q) \right|^2} = O\left( (\Gamma_M)^{-1}(\Gamma_M^{[\varsigma/2]})^2\right).\] So it remains to consider the $N_M$ term. Note that since the $U_k^q$ are independent standard Gaussian vectors, \((C_0 \sqrt{\Gamma_M})^{-1} N_{F^i,M}(x,q)\) when $i$ ranges $1,\ldots,d_x$ is a Gaussian vector. Let us study its covariance matrix $\Phi_F^{n}$. Using \eqref{Eq:EstimateProductCentered} we get for $i,j=1,\ldots,n$ \[\begin{split} &\Phi_F^{i,j;n}(x,q):= \EX{ \frac{1}{\Gamma_M} N_{F^i,M}(x,q) N_{F^j,M}(x,q') }\\ &=\one{q=q'}\sum_{k,k'=1}^{M} \gamma_k \langle \sigma^*(\cdot)D_y\phi_{F^i}(\cdot), \sigma^*(\cdot)D_y\phi_{F^{j}}(\cdot) \rangle (x,\DecrAvg{Y}^{x,q}_{k-1}). \end{split} \] Define $\varphi^n_F=\sqrt{\Phi^n_F}$ (the Cholesky decomposition). Then, there exists a family of independent gaussian variables $\nu_{t_k}^{i,j;n}$, $\bar{\ensuremath \mathcal{F}}_{t_k}$-measurable such that $ (\Gamma_M)^{-1} N_{F^i,M}(x,q) = \sum_{j=1}^{d_x} \varphi^{i,j;n}\nu_{t_k}^{i,j;n}$. Moreover from Proposition \ref{Prop:ConvergenceEstimator} and Proposition \ref{Prop:ControlDecreasingStep} we have that $\Phi^n_F(x,q) $ converges uniformly in $x$ in $L^2$ to $\Phi_F(x)$ as defined in the claim with rate $O(n^{-\beta})$. By Theorem \ref{Theo:ControlCholesky} we get the same uniform convergence for $\varphi_F^n$. The claim follows in this case. \item The case $\beta_0 = \beta_1$ is straightforward from what has been proven in the previous cases. \end{itemize} \item Since $H, \DecrAvg{H}^n$ satisfy the same properties as $F,\DecrAvg{F}^n$, we get the claim for $R_H$, $\varphi_H$ and $\nu_k^{i,j;n}$ by analogous arguments. Replacing this result in the sensitivity of the Cholesky procedure given in Lemma \ref{Lem:ErrorCholesky}, and taking into account the independence of the Gaussian entries, we get the claim for $R_G$ and $\varphi_G$. \end{enumerate} \end{proof} Let $\{\upsilon_n\}$ be a sequence of increasing positive numbers and let us consider the sequence of rescaled error processes $\zeta^{n}$, defined by \[\zeta_t^{n}:= \upsilon_n (X_t - \ApproxVar{X}^{n}_t ).\] We can show that this sequence of processes converges in distribution in the uniform convergence topology to a process $\zeta$ defined as the solution to a certain stochastic differential equation. We divide the analysis in two main cases: a first one in which $G(x)\equiv 0 $, i.e. when $X$ is the solution to an ordinary differential equation; and the case when $G(x)$ is non degenerate. Just as in the asymptotic error obtained for the usual stochastic Euler method given in \cite{jacod_asymptotic_1998}, we will obtain different rates and different components in the equation for both cases. \begin{MyTheorem}[Limit distribution] \label{Theo:AbstractLimitDistribution} Under the assumptions and notation of Proposition \ref{Prop:MainDecreasingStepPqCLT}, let $\rho,R_F,\varphi_F,R_G,\varphi_G$ be defined as in Proposition \ref{Prop:MainDecreasingStepPqCLT} and $\beta$ defined in \eqref{Eq:DefinitionBeta}. \begin{enumerate}[i)] \item \label{Enum:AbstractTheoLimitDistODE} (ODE case- $G(x)\equiv 0$.) Let $B^{1}$ be the Brownian process given in Proposition \ref{Prop:ConvergenceOfTuple}. Let $r = 1 \wedge (1/2 + \beta)$, and suppose $\rho\geq r-\beta$. Let \[\zeta_t^{n}:= n^r (X_t - \ApproxVar{X}^{n}_t ).\] Then $\zeta^n\Rightarrow \zeta^\infty$ in the uniform convergence sense, where $\zeta^\infty$ is solution of the system \begin{multline} \zeta_t^{\infty,i} = \sum_{j=1}^{d_x} \left( \int_0^t \partial_{x^j} F^i(X_s)\zeta^{\infty,j}_s ds + \one{\beta\geq 1/2} \frac{1}{2} \int_0^t \partial_{x^j} F^i(X_s) F^j(X_s) ds \right) \\ + \one{\beta\leq 1/2} \left( \int_0^t R_{F}^{i}(X_s) ds + \sum_{l=1}^{d_x}\int_0^t \varphi_F^{i,l}(X_s) dB^{1;l}_s\right).\label{Eq:LimitDistODE} \end{multline} \item \label{Enum:AbstractTheoLimitDistSDE} (SDE case - $G(x)\neq 0$) Let $B^{2}$ and $B^{3}$ be the independent Brownian processes given in Proposition \ref{Prop:ConvergenceOfTuple}. Let $r = (1/2 \wedge \beta )$ and \[\zeta_t^{n}:= n^r (X_t - \ApproxVar{X}^{n}_t ).\] Then $\zeta^n \Rightarrow \zeta^\infty$ , where $\zeta^\infty$ is solution of the system for $i=1,\ldots,d_x$ of \begin{equation} \begin{split} \zeta_t^{\infty,i} = & \sum_{j} \left( \int_0^t \partial_{x^j} F^i(X_s)\zeta^{\infty,j}_s ds + \int_0^t R_G^{i,j}(X_s) dW^j_s \right)\\ & + \one{\beta \leq 1/2} \sum_{j,k,l =1}^{d_x}\int_0^t \varphi_G^{i,j,l,k}(X_s) dB^{3;l,k,j}_s\\ & + \one{\beta \leq 1/2} \sum_{j,l=1}^{d_x} \int_0^t \partial_{x^j} G^{i,l}(X_s)\zeta^{\infty,j}_s dW^l_s \\ & + \one{\beta \geq 1/2} \frac{1}{\sqrt{2}} \sum_{j,k,l=1}^{d_x} \int_0^t \partial_{x^j} G^{i,l}(X_s) G^{j,k}(X_s) dB^{2;k,l}_s. \end{split} \label{Eq:limiDistSDE} \end{equation} \end{enumerate} \end{MyTheorem} Let us remark that if $\beta > 1/2$ in Theorem \ref{Theo:AbstractLimitDistribution}, the error of the Euler scheme dominates: we recover the limit distribution error for an Euler scheme with exact coefficients given in \cite{kurtz_wong-zakai_1991} or \cite{jacod_asymptotic_1998}. By contrast, if $\beta<1/2$, it is the decreasing Euler estimate error that becomes dominant. Since a higher $\beta$ is general only achieved by paying a higher price in the required number of steps for the decreasing Euler step, the optimal choice implies fixing $\beta=1/2$. Before proving Theorem \ref{Theo:AbstractLimitDistribution}, let us show how it implies Theorem \ref{Theo:ShortMainDecreasingStep} \begin{proof}[Proof of Theorem \ref{Theo:ShortMainDecreasingStep}] The result is obtained, from Theorems \ref{Theo:GeneralConvergence} and \ref{Theo:AbstractLimitDistribution}, since $\ensuremath {(\mathcal{H}_{s.s.})}$ and $\ensuremath {(\mathcal{H}_{f.s.})}$ are directly assumed and as the sequence defined as $\gamma_k = \gamma_1 k^{-\theta}$ for $0<\theta<1$ satisfies Hypothesis $\ensuremath {(\mathcal{H}_{\gamma})}$. Moreover, recall that we fixed $M(n)= \lceil M_1n^{1/(1-\theta)} \rceil $, and we have for $n$ large enough \begin{gather*} \Gamma_M \approx \frac{\gamma_0 M_1^{1-\theta}n}{1-\theta}, \quad\quad \frac{\Gamma_M^{[\varsigma/2]}}{\Gamma_M} \approx \frac{(1-\theta) M_1^{-(\varsigma/2-1)\theta} n^{-\frac{(\varsigma/2-1)\theta}{1-\theta}}}{1-\varsigma\theta/2},\\ \frac{\Gamma_M^{[\varsigma/2+1]}}{\Gamma_M^{[\varsigma/2]}} \approx \frac{ (1-\varsigma\theta/2) M_1^{-\theta} n^{-\frac{\theta}{1-\theta}}}{1-(\varsigma/2+1)\theta}; \end{gather*} So that we get from Proposition \ref{Prop:MainDecreasingStepPqCLT}, that $\beta_0 = 1/2$ and \begin{align*} \beta_1&=\frac{(\varsigma/2-1)\theta}{1-\theta}, & \beta_2&=\frac{\theta}{1-\theta}, & C_0 &\approx \frac{\gamma_0 M_1^{1-\theta}}{1-\theta}, & C_1 &\approx \frac{(1-\theta) M_1^{-\theta}}{1-2\theta}. \end{align*} Recall that $\varsigma$ is defined in \eqref{Eq:DefinitionLbar} and stands for the first non-zero term in the error expansion of the decreasing Euler estimator. Let us assume we are in the worst case when it attains its minimal value $\varsigma=4$. Hence \[\beta_1 = \frac{\theta}{1-\theta},\quad\quad C_1 = \frac{(1-\theta) M_1^{-\theta}}{1-2\theta}.\] Let us now deduce The conditions on $\theta$ are then deduced from the conditions in Theorem \ref{Theo:AbstractLimitDistribution} for each of our study cases: \begin{itemize} \item ODE with random coefficients: From the conditions of Theorem \ref{Theo:AbstractLimitDistribution} we have \[r = 1 \wedge \left(\frac{1}{2}+\beta \right) = \frac{1}{2} + \left(\frac{1}{2}\wedge \beta \right) = \frac{1}{2} + \left(\frac{1}{2}\wedge (\beta_0 \wedge \beta_1)\right) = \frac{1}{2} + \beta \] since we should verify $\rho \geq r-\beta = 1/2 $, this implies \[ |\beta_0 -\beta_1| = \left|\frac{1}{2}-\frac{\theta}{1-\theta}\right|\geq \rho \geq \frac{1}{2},\] which is the case if $\theta \in [1/2,1)$. Moreover, since in this case $\beta_1\geq1>\beta_0=1/2$, we get $r=1/2$, and the $R_F$ term disappears. \item Full SDE case: We have $r =\beta = 1/2 \wedge (\theta/(1-\theta))$ the only restriction comes from imposing $\beta=1/2$. This is obtained for $1/3\leq \theta<1$. Note that the $R_G$ term is different from zero only if $\theta=1/3$. \end{itemize} Finally, note that if $\varsigma> 4$, we get from the constraints $\theta\in [1/2,1)$ in the ODE with random coefficients case that $\beta_1> \beta_0+1/2$ and from from fixing $\theta \in [1/3,1)$ in the full SDE case that $\beta_1>\beta_0=1/2$, $\beta_1>\beta_2$. In both those cases the $R_G$ term is zero. \end{proof} \begin{MyRemark} It should be noted from the proof of Theorem \ref{Theo:ShortMainDecreasingStep} that knowing a priori that $\varsigma>4$ makes it possible to obtain a lower inferior bound for $\theta$ in the theorem. Since in general we do not know $\varsigma$, we have stated our results with the sometimes sub-optimal limits. \end{MyRemark} \begin{proof}[Proof of Theorem \ref{Theo:AbstractLimitDistribution}] \begin{enumerate}[a)] \item Let us deal first with the full SDE case. We have from the definition of $\zeta^{n}$ that \begin{equation} \zeta^{n}_t = \int_0^t n^{r} \left( F(X_s) - \DecrAvg{F}^{n}(\ApproxVar{X}^{n}_{\tbar[s]},\tbar[s]) \right) ds + \int_0^t n^{r} \left( G(X_s) - \DecrAvg{G}^{n}(\ApproxVar{X}^{n}_{\tbar[s]},\tbar[s]) \right) dW_s.\label{Eq:developmentZita} \end{equation} Let us examine each one of these terms separately. Denoting by $x^i$ the $i-th$ component of $x$, let $x,y \in \ensuremath \mathds{R}^{d_x}$. We define the set of vectors $\Delta^j (x,y)$ \[\Delta^j(x,y) := \begin{cases} x & \text{for } j=0\\ (y_1,y_2,\ldots,y_j,x_{j+1},x_{j+2},\ldots ,x_{d_x})^* & \text{for } 1\leq j\leq d_x\end{cases}\] and \[\Delta^j F^i (x,y) := \one{ x^j \neq y^j } \left( \frac { F^i(\Delta^{j-1}(x,y)) - F^i(\Delta^j(x,y)) }{ x^j - y^j } \right) + \one{ x^j = y^j } \partial_{x^j}F^i(x),\] and recalling that \[ \ApproxVar{X}^{j,n}_{s} - \ApproxVar{X}^{j,n}_{\tbar[s]} = F^j(\ApproxVar{X}^{n}_{\tbar[s]})(s-\tbar[s]) + \sum_{l=1}^{d_x} G^{j,l}(\ApproxVar{X}^{n}_{\tbar[s]})(W^l_s-W^l_{\tbar[s]}), \] we have \begin{align*} \int_0^t & n^{r} \left[ F^i(X_s) - \DecrAvg{F}^{i;n}(\ApproxVar{X}^{n}_{\tbar[s]},\tbar[s]) \right] ds \\ & = \int_0^t n^{r} ( F^i(X_s) - F^i(\ApproxVar{X}^{n}_{s}) ) ds + \int_0^t n^{r} \left( F^i(\ApproxVar{X}^{n}_{s}) - F^i(\ApproxVar{X}^{n}_{\tbar[s]}) \right) ds +\\ & \quad\quad + \int_0^t n^{r} \left( F^i(\ApproxVar{X}^{n}_{\tbar[s]}) - \DecrAvg{F}^{i;n}(\ApproxVar{X}^{n}_{\tbar[s]},\tbar[s]) \right) ds\\ \end{align*} \begin{align*} % & = \int_0^t \sum\limits_j \left[ \vphantom{ \frac{1}{x^2} } n^{r} \Delta^j F^i (X_s, \ApproxVar{X}^{n}_{s} ) (X^j_s - \ApproxVar{X}^{j;n}_{s}) \right. \\ & \quad\quad\quad +n^{r}\Delta^j F^i (\ApproxVar{X}^{n}_s, \ApproxVar{X}^{n}_{\tbar[s]} ) F^j(\ApproxVar{X}^{n}_{\tbar[s]}) (s-\tbar[s]) \\ & \quad\quad\quad \left. +\sum_{l=1}^{d_x}n^{r}\Delta^j F^i (\ApproxVar{X}^{n}_s, \ApproxVar{X}^{n}_{\tbar[s]} ) G^{j,l}(\ApproxVar{X}^{n}_{\tbar[s]}) (W^l_s-W^l_{\tbar[s]}) \right]ds \\ & \quad + n^{r-\beta}\int_0^t n^{\beta}\left( F^i(\ApproxVar{X}^{n}_{\tbar[s]}) - \DecrAvg{F}^{i;n}(\ApproxVar{X}^{n}_{\tbar[s]},\tbar[s]) \right) ds. \end{align*} Following the same approach we obtain for each $l=1\ldots,d_x$ \begin{align*} \int_0^t & n^{r} \left[ G^{i,l}(X_s) - \DecrAvg{G}^{i,l;n}(\ApproxVar{X}^{n}_{\tbar[s]},\tbar[s]) \right] dW^l_s.\\ & = \int_0^t \sum\limits_j \left[ \vphantom{ \frac{1}{x^2} } n^{r} \Delta^j G^{i,l} (X_s, \ApproxVar{X}^{n}_{s} ) (X^j_s - \ApproxVar{X}^{j;n}_{s}) \right. \notag\\ & \quad\quad\quad +n^{r}\Delta^j G^{i,l} (\ApproxVar{X}^{n}_s, \ApproxVar{X}^{n}_{\tbar[s]} ) F^j(\ApproxVar{X}^{n}_{\tbar[s]}) (s-\tbar[s]) \notag\\ & \quad\quad\quad \left.+\sum_{k=1}^{d_x}n^{r}\Delta^j G^{i,l} (\ApproxVar{X}^{n}_s, \ApproxVar{X}^{n}_{\tbar[s]} ) G^{j,k}(\ApproxVar{X}^{n}_{\tbar[s]}) (W^k_s-W^k_{\tbar[s]}) \right]dW^l_s\\ & \quad + \vphantom{ \frac{1}{x^2} } n^{r-\beta}\int_0^t n^{\beta}\left( G^{i,l}(\ApproxVar{X}^{n}_{\tbar[s]}) - \DecrAvg{G}^{i,l;n}(\ApproxVar{X}^{n}_{\tbar[s]},\tbar[s]) \right) dW^l_s.\notag \end{align*} By identifying terms in the obvious way, we write \[ \zeta_t^{i,n} = ( P^{i,n}_1(t)+P^{i,n}_2(t)) + \int_0^t \langle Q^{i;n}_1(s), \zeta_s^{n} \rangle ds + \sum_{l=1}^{d_x}\int_0^t \langle Q^{i,l;n}_2(s), \zeta_s^{n} \rangle dW^l_s.\] where $Q_1^i$, $Q_2^{i,l}$ are $d_x$ dimensional random processes with components \begin{align*} Q^{j,i;n}_1(s) &= \Delta^j F^i (X_s, \ApproxVar{X}^{n}_{s}) \quad\quad Q^{j,i,l;n}_2(s) = \Delta^j G^{i,l} (X_s, \ApproxVar{X}^{n}_{s}) \end{align*} and \begin{align*} P^{i;n}_2(s) = & n^{r-\beta}\int_0^t n^{\beta}\left( F^i(\ApproxVar{X}^{n}_{\tbar[s]}) - \DecrAvg{F}^{i;n}(\ApproxVar{X}^{n}_{\tbar[s]},\tbar[s])\right)ds\\ & \quad + n^{r-\beta}\int_0^t n^{\beta}\left( G^{i,l}(\ApproxVar{X}^{n}_{\tbar[s]}) - \DecrAvg{G}^{i,l;n}(\ApproxVar{X}^{n}_{\tbar[s]},\tbar[s]) \right) dW^l_s. \end{align*} \begin{align*} P^{i;n}_1(s) = & \int_0^t \sum\limits_j \left[ \vphantom{\sum^{d_x}}n^{r}\Delta^j F^i (\ApproxVar{X}^{n}_s, \ApproxVar{X}^{n}_{\tbar[s]} ) F^j(\ApproxVar{X}^{n}_{\tbar[s]}) (s-\tbar[s]) \right.\\ & \quad\quad +\left.\sum_{l=1}^{d_x}n^{r}\Delta^j F^i (\ApproxVar{X}^{n}_s, \ApproxVar{X}^{n}_{\tbar[s]} ) G^{j,l}(\ApproxVar{X}^{n}_{\tbar[s]}) (W^l_s-W^l_{\tbar[s]}) \right]ds \\ &\quad+ \int_0^t \sum\limits_{j,l=1}^{d_x} \left[ \vphantom{\sum^{d_x}} n^{r}\Delta^j G^{i,l} (\ApproxVar{X}^{n}_s, \ApproxVar{X}^{n}_{\tbar[s]} ) F^j(\ApproxVar{X}^{n}_{\tbar[s]}) (s-\tbar[s]) dW^l_s \right. \\ & \quad\quad +\left.\sum_{k=1}^{d_x}n^{r}\Delta^j G^{i,l} (\ApproxVar{X}^{n}_s, \ApproxVar{X}^{n}_{\tbar[s]} ) G^{j,k}(\ApproxVar{X}^{n}_{\tbar[s]}) (W^k_s-W^k_{\tbar[s]})\right] dW^l_s \end{align*} \item In this step, we introduce a nicer diffusion and study its convergence, and prove it shares the limit distribution of the previous SDE. Let \[ \check{\zeta}_t^{i,n} = ( \check{P}^{i,n}_1(t)+\check{P}^{i,n}_2(t)) + \int_0^t \langle \check{Q}^{i;n}_1(s), \check{\zeta}_s^{n} \rangle ds + \sum_{l=1}^{d_x}\int_0^t \langle \check{Q}^{i,l;n}_2(s), \check{\zeta}_s^{n} \rangle dW^l_s.\] where \begin{align*} \check{Q}^{i;n}_1(s) = & \nabla F^{i}(X_s) ; \quad \quad\quad \check{Q}^{i,l;n}_2(s) = \nabla G^{i,l}(X_s) ; \\ \check{P}^{i;n}_1(s) = & \frac{1}{2}\int_0^t n^r \left\langle \nabla F^i (X_s), F(X_s) \right\rangle dA^{0;n} \\ & + \sum_{l=1}^{d_x}\int_0^t n^{r} \langle \nabla F^i (X_s ), G^{\cdot,l}(X_s)\rangle dA_s^{1;l;n} \\ &+ \int_0^t n^{r} \langle \nabla G^{i,l} (X_s), F(X_s)\rangle dB_s^{0;l,n} \\ &+ \sum_{k,l=1}^{d_x} \frac{1}{\sqrt{2}} \int_0^t n^{r} \langle \nabla G^{i,l} (X_s) G^{\cdot,k}(X_s) \rangle dB_{s}^{2;k,l,n}, \end{align*} \begin{align*} \check{P}^{i;n}_2(s) = & n^{r-\beta}\int_0^t \sum_{j,k,l=1}^{d_x}\varphi_{G}^{i,j,l,k}(X_s) dB_s^{3;l,k,j,n}+n^{r-\beta} \int_0^t \sum_{j=1}^{d_x} R_{G}^{i,j}(X_s) dW^j_s\\ & + n^{r-\beta}\int_0^t \sum_{j=1}^{d_x}\varphi_{F}^{i,j}(X_s) dB_s^{1;j,n} + n^{r-\beta} \int_0^t R_{F}^{i}(X_s) ds, \end{align*} for $R_F,R_G,\varphi_F,\varphi_G$ defined in Proposition \ref{Prop:MainDecreasingStepPqCLT}. By $\ensuremath {(\mathcal{H}_{s.s.})}$, $F,G$ are bounded; by Lemma \ref{Lem:PropertiesAverage}, $\nabla F$ and $\nabla G$ are well defined, bounded, and have bounded derivatives; and from the definition of $R_F,R_G,\varphi_F,\varphi_G$ are $C_b^1$. Note that \eqref{Eq:TupleConv3} in Proposition \ref{Prop:ConvergenceOfTuple} gives us goodness and convergence of the tuple $(n^r A^{0,n},n^r A^{1,n},n^r B^{0,n},B^{1,n},n^r B^{2,n}, B^{3,n})$. Hence, by virtue of Theorem 5.4 in \cite{kurtz_weak_1991} $\check{\zeta}^{n}(\cdot \wedge \tau_K^{n})$ is tight and any limit point will satisfy \eqref{Eq:limiDistSDE} on the interval $[0,\tau_K]$ where $\tau_K = (\inf\{t: |\zeta(t)|>K\} \wedge T)$. Moreover \begin{align*} \sup_{0\leq s \leq \tau_K } \|\check{P}^{n}_s\| &, &\sup_{0\leq s \leq \tau_K } \|\check{Q}_1^{n}(s)\| &, &\sup_{0\leq s \leq \tau_K } \|\check{Q}_2^{n}(s)\| &; \end{align*} are tight. \item We prove now that both $\zeta^n$ and $\check{\zeta}^n$ have the same limit on the interval $[0,\tau_K]$. By Theorem \ref{Thm:ConvergenceChangeCoefficients}, it suffices to prove that sup norm of the difference of the coefficients converge in probability. By Theorem \ref{Theo:GeneralConvergence} the regularity properties of $F$ and the mean value theorem we have \[\EX{\sup_{0\leq t \leq \tau_K} | Q_1^{i,n}(t) - \check{Q}_1^{i,n}(t)|} \leq \EX{ \sup_{x} | D^2 F(x) | \sup_{0\leq t \leq \tau_K} | X_t - \ApproxVar{X}_t^n|} \rightarrow 0.\] The terms of $Q_2^{n},P_1^{n}$ are treated in the same way. On the other hand, we get from Corollary \ref{Cor:ControlPowerAndDifference}, Proposition \ref{Prop:APriori}, and Burckholder-Davis-Gundy inequality that \begin{gather*} \sup_{0\leq t\leq T}\left| n^{r-\beta}\int_0^t \left[ n^{\beta}\left( F^i(\cdot) - \DecrAvg{F}^{i;n}(\cdot, \tbar[s]) \right) - \sum_{j=1}^{d_x}\varphi_F^{i,j}(\cdot) \nu^{j;n}_{\tbar[s]} - R_{F}^i(\cdot) \right](\ApproxVar{X}^{n}_{\tbar[s]}) ds, \right|\\ \begin{split} \sup_{0\leq t\leq T} & \left| n^{r-\beta}\int_0^t \left[ n^{\beta}\left( G^{i,j}(\cdot) - \DecrAvg{G}^{i,j;n}(\cdot,\tbar[s]) \right) \right.\right.\\ & \quad\quad\quad\quad\quad\quad\quad - \left.\left.\sum_{j,k=1}^{d_x}\varphi_G^{i,j,l,k}(\cdot) \nu^{l,k;n}_{\tbar[s]} - R_{G}^{i,j}(\cdot) \right](\ApproxVar{X}^{n}_{\tbar[s]}) dW^l_s \right| \end{split} \end{gather*} are tight and converge to zero. Thus, by Theorem \ref{Thm:ConvergenceChangeCoefficients} we will have that $\zeta^{i;n}$ and $\check{\zeta}^{i;n}$ will converge to the same limit. \item Finally, note that $\tau^{n}_K\rightarrow \infty$ and $\tau_K \rightarrow \infty$, proving our claim in the full SDE case. \item To prove $\ref{Enum:AbstractTheoLimitDistODE})$ it suffices to follow the same approach. We obtain an equivalent development for the ODE with random coefficients case (replacing by zero all the ``g-terms''). The rest of the proof proceeds as before, this time using \eqref{Eq:TupleConv1} for the weak convergence of the tuple. \end{enumerate} \end{proof} \section{The EMsDS algorithm} \label{Sec:EMSDSalgorithm} Given the error expansion for the decreasing step algorithm presented in Proposition \ref{Prop:ErrorExpansion}, it seems natural to explore if a Richardson-Romberg extrapolation may be used to obtain the approximation with the same convergence properties we have proven. The idea of such a procedure is to decrease the complexity by performing a linear combination of two (or more) realizations of the algorithm with carefully chosen parameters. We borrow here the procedure as defined in \cite{lemaire_estimation_2005}. Let $\lambda$ be a positive real. If $\{\gamma_k\}$ is a sequence of steps satisfying $\mathcal{H}_{\lambda}$, the sequence $\gamma_k^\lambda:= \frac{\gamma_k}{\lambda}$ will also satisfy $(\mathcal{H}_{\lambda})$. We will denote $\Gamma^\lambda_M$ and $\Gamma^{\lambda,[r]}_M$ the sum of the $\gamma_k^\lambda$ and its power as before. Let us denote by $\DecrAvg{F}^{\lambda,M}(x,q)$ the approximation as defined in \eqref{Eq:AverageEstimator} when the coefficients $\{\gamma_k^\lambda\}_{k\in \ensuremath \mathds{N}^*}$ are used. With $\varsigma$ be given as in \eqref{Eq:DefinitionLbar}, let us define the \emph{extrapolated approximation estimator} as \begin{equation} \label{Eq:DefinitionFHat} \DecrAvgExtrap{F}^{\lambda;M(n)}(x,q) = \frac{1}{\lambda^{\varsigma/2-1}-1} \left(\lambda^{\varsigma/2-1} \DecrAvg{F}^{\lambda, M(n) }(x,q) - \DecrAvg{F}^{ M(n) }(x,q) \right) \end{equation} The first question we might ask is if estimator \eqref{Eq:DefinitionFHat} does converge to the actual ergodic average, and what type of properties it inherits. To clarify the situation consider an extension of \eqref{Eq:ErgodicDiffusion}. Let $\vec{Y}^x=(Y^{1;x},Y^{2;x})^*$ with \begin{equation}\label{Eq:ExtendedSystem} \begin{split} Y_{t}^{1;x} = y_0^1 + \int_0^t \frac{b(x,Y_s^{1;x})}{\lambda} ds + \int_0^t \frac{\sigma (x, Y_s^{1;x})}{\sqrt{\lambda}} d\ApproxVarExtrap{W}^1_s,\\ Y_t^{2;x} = y_0^2 + \int_0^t b(x,Y_s^{2;x}) ds + \int_0^t \sigma (x, Y_s^{2;x}) d\ApproxVarExtrap{W}^2_s. \end{split} \end{equation} If $\ApproxVarExtrap{W}^1$ and $\ApproxVarExtrap{W}^2$ are independent, then this system satisfies $\ensuremath {(\mathcal{H}_{f.s.})}$ with a unique invariant measure defined by $\vec{\mu}^x(d\vec{y}) = \mu^x(dy^1)\mu^x(dy^2)$. If we define \begin{equation} \label{Eq:ExtendedF} \vec{f}(x,\vec{y}) := \frac{1}{\lambda^{\varsigma/2-1}-1}\left(\lambda^{\varsigma/2-1} f(x,y^1)-f(x,y^2)\right), \end{equation} and defining in an analogous way $\vec{h}$, then it can be seen that $\vec{f}, \vec{g}, \vec{h}:=\vec{g}\vec{g}^*$ satisfy $\ensuremath {(\mathcal{H}_{s.s.})}$. Moreover if we apply the decreasing step algorithm to $\vec{f}$ (respectively $\vec{h}$) in the extended framework, we obtain the expression \eqref{Eq:DefinitionFHat}. \emph{Hence, we conclude that the EMsDS algorithm is equivalent to the MsDS algorithm applied to an extended system.} Let us denote by $\ApproxVarExtrap{X}^{n}$ the approximation of the diffusion $X$ using the extrapolated version of the algorithm. In view of the discussion we presented before, the following result is mainly a corollary of Theorems \ref{Theo:GeneralConvergence} and \ref{Theo:AbstractLimitDistribution}, and extends the main Theorem to the extrapolation algorithm. It shows the advantage of using the EMsDS algorithm: assuming higher regularity, all the properties of the MsDS algorithm are conserved but the extrapolated version \emph{allows a lower value for $\theta$ in the definition of the sequence $\gamma_k = \gamma_0 k^{-\theta}$}. More precisely we pass from $1/2$ to $1/3$ in the ODE case and from $1/3$ to $1/5$ in the SDE case as minimal $\theta$ values. As a consequence of this reduction, the complexity of the modified version is in general asymptotically lower than that of the non extrapolated version (refer to the efficiency analysis on subsection \ref{Subsec:Efficiency}). \begin{MyTheorem} \label{Theo:ShortMainDecreasingStepInterp} Let $0<\theta<1$, $\gamma_1 \in \ensuremath \mathds{R}^+$ and $\gamma_k = \gamma_1 k^{-\theta}$. Assume $\ensuremath {(\mathcal{H}_{f.s.})}$ and $\ensuremath {(\mathcal{H}_{s.s.})}$, $M(n)$ defined as in Theorem \ref{Theo:ShortMainDecreasingStep}, and assume in addition that $r_y> 5$. Let $\ApproxVarExtrap{X}^n$ be the approximated diffusion where we replace the ergodic estimator \eqref{Eq:AverageEstimator} by \eqref{Eq:DefinitionFHat}. \begin{enumerate}[i)] \item (Strong convergence). There exists a constant $K$ such that \begin{itemize} \item Case $g\equiv 0$ (ODE with random coefficients): \[\EX{\sup\limits_{0\leq t \leq T}\left|X_t - \ApproxVarExtrap{X}_t^{n}\right|^2 } \leq K n^{-2[(1-\theta)\wedge 2\theta]/(1-\theta)}.\] \item (Full SDE case): \[\EX{\sup\limits_{0\leq t \leq T}\left|X_t - \ApproxVarExtrap{X}_t^{n}\right|^2 } \leq K n^{-[(1-\theta)\wedge 4\theta]/(1-\theta)}.\] \end{itemize} \item (Limit distribution). Assume in addition that $r^y \geq 8$, and define \begin{align*} \ApproxVarExtrap{C}_\varphi &:= \frac{(\lambda^{3}+1)^{1/2}}{\lambda-1}; \quad\quad\quad\quad\quad \ApproxVarExtrap{C}_1 := \frac{\gamma_0^2 (1-\theta) M_1^{-2\theta} }{1-3\theta};\\ \ApproxVarExtrap{\varphi}_F(x) &:= \one{\theta = 1/5}\ApproxVarExtrap{C}_\varphi \sqrt{\ApproxVarExtrap{\Phi}_F(x)}; \quad \ApproxVarExtrap{\varphi}_G(x) := \one{\theta = 1/5}\ApproxVarExtrap{C}_\varphi \sqrt{\ApproxVarExtrap{\Phi}_G(x)}; \\ \ApproxVarExtrap{R}_{F}^{i}(x) &:= \one{\theta\geq 1/5} \ApproxVarExtrap{C}_1(1-\lambda^{-1}) \int \bar{v}_{F^i}^{\varsigma+2,r^y}(x,y) \mu^x(dy); \\ \ApproxVarExtrap{R}_{H}^{i,j}(x) &:= \one{\theta\geq 1/5} \ApproxVarExtrap{C}_1(1-\lambda^{-1})\int \bar{v}_{H^{i,j}}^{\varsigma+2,r^y}(x,y) \mu^x(dy). \end{align*} \begin{itemize} \item \label{Enum:TheoLimitDistInterpODE} (ODE case: $G(x)\equiv 0$). If $\theta \geq 1/3 $, then \(\ApproxVarExtrap{\zeta}^{n}:= n \left(X_t - \ApproxVarExtrap{X}^{n} \right) \) satisfies the limit distribution result given in Theorem \ref{Theo:ShortMainDecreasingStep} $\ref{Enum:TheoLimitDistODE})$ with new coefficients $ \ApproxVarExtrap{\varphi}^F$ instead of $\varphi^F$. \item \label{Enum:TheoLimitDistInterpSDE} (SDE case)If $\theta \geq 1/5$, then \(\ApproxVarExtrap{\zeta}^{n}:= n^{1/2} \left(X_t - \ApproxVarExtrap{X}^{n} \right) \) satisfies the limit distribution result given in Theorem \ref{Theo:ShortMainDecreasingStep} $\ref{Enum:TheoLimitDistSDE})$ with the coefficients $\ApproxVarExtrap{R}^F,\ApproxVarExtrap{R}^G,\ApproxVarExtrap{\varphi}^F$ and $\ApproxVarExtrap{\varphi}^G$ instead of $R^F$, $R^G$, $\varphi_F$ and $\varphi_g$ respectively. \end{itemize} \end{enumerate} \end{MyTheorem} \begin{proof}[Proof of Theorem \ref{Theo:ShortMainDecreasingStepInterp}] We will deduce the proof only for the full SDE case the other case being analogous. We assume that $\varsigma=4$, which is the most common case. \begin{enumerate}[a)] \item As in the proof of Theorem \ref{Theo:ShortMainDecreasingStep}, the sequence of coefficients satisfies $\ensuremath {(\mathcal{H}_{\gamma})}$. Moreover, the EMsDS algorithm is the MsDS algorithm applied to an extended system, and hence the strong convergence and limit distribution properties are a consequence from Theorems \ref{Theo:GeneralConvergence} and \ref{Theo:AbstractLimitDistribution}: it remains just to express the values of the functions and constants appearing in Propositions \ref{Prop:ControlDecreasingStep} and \ref{Prop:MainDecreasingStepPqCLT} in terms of the original system. Indeed, recall that \begin{align} \label{Eq:DefinitionDerivativeHat} \vec{b}(x,\vec{y}) & = \begin{pmatrix} \lambda^{-1}b(x,y^1) \\ b(x,y^2)\end{pmatrix}; & \vec{\sigma}(x,\vec{y}) & = \begin{pmatrix} \lambda^{-1/2}\sigma(x,y^1) & 0 \\0 & \sigma(x,y^2) \end{pmatrix}; \end{align} By $i)$ in Proposition \ref{Prop:ControlDecreasingStep} applied to the extended problem (i.e for the system \eqref{Eq:ExtendedSystem} and $\vec{f}$ defined in \eqref{Eq:ExtendedF}), we have a solution for the extended centered Poisson equation given by \[\vec{\phi}_{F^i}(x,\vec{y}) = (\lambda-1)^{-1} \left(\lambda^{2}\phi_{F^i}(x,y^1) - \phi_{F^i}(x,y^2)\right),\] i.e., the solution of equation \eqref{Eq:CenteredPoisson} with function $F^i$ under the extended set-up is a linear combination of the solution in the original set-up. Thus, for any $j>0$, \begin{equation} \label{Eq:DerivativePhiHat} D^j_y \vec{\phi}_{F^i}(x,\vec{y}) = \frac{1}{\lambda-1} \begin{pmatrix} \lambda^{2} D^j_y\phi_{F^i}(x,y^1)\\ - D^j_y \phi_{F^i}(x,y^2)\end{pmatrix}. \end{equation} It follows that \begin{align*} & D^j_y\vec{\phi}_{F^i}(x,\vec{y})\EX{\left\langle \vec{b}(x,\vec{y})^{\otimes(l-j)}, (\vec{\sigma}(x,\vec{y}) U_1^0)^{\otimes(2j-l)}\right\rangle}\\ & = \frac{\lambda^{2}}{\lambda-1} D^j_y\phi_{F^i}(x,y^1)\EX{ \left\langle\left(\frac{b(x,y^1)}{\lambda}\right)^{\otimes(l-j)}, \left(\frac{\sigma(x,y^1)}{\sqrt{\lambda}} U_1^0\right)^{\otimes(2j-l)}\right\rangle}\\ & \quad -\frac{1}{\lambda-1} D^j_y\phi_{F^i}(x,y^2)\EX{\left\langle \vec{b}(x,y^2)^{\otimes(l-j)}, (\sigma(x,y^2) U_1^0)^{\otimes(2j-l)}\right\rangle}. \end{align*} Therefore \begin{equation} \label{Eq:v_arrow_bar} \vec{\bar{v}}^{l,r^y}_{F^i} =\left(\frac{\lambda^{(4-l)/2}-1}{\lambda-1} \right)\bar{v}^{l,r^y}_{F^i}, \end{equation} and we deduce that the terms of the error expansion will be zero for $l\leq 5$. \item Let $\vec{\varsigma}$ be defined by \eqref{Eq:DefinitionLbar} under the extended setup. From \eqref{Eq:v_arrow_bar} we conclude that $\vec{\varsigma}\geq 6$, being $\vec{\varsigma}=6$ the worst case. Hence, we deduce that defining \begin{align*} \beta_0 &= \frac{1}{2}, &\ApproxVarExtrap{\beta_1} & =\frac{2\theta}{1-\theta} &\hat{\beta}_2 &=\frac{\theta}{1-\theta}. \end{align*} then, \begin{align*} \Gamma_M & \approx \frac{\gamma_0 M_1^{1-\theta}n}{1-\theta}, & \frac{\Gamma_M^{[3]}}{\Gamma_M} & \approx \frac{\gamma_0^2 (1-\theta) M_1^{-2\theta} n^{-\frac{2\theta}{1-\theta}}}{1-3\theta},\\ \frac{\Gamma_M^{[4]}}{\Gamma_M^{[3]}} & \approx \frac{\gamma_0 (1-3\theta) M_1^{-\theta} n^{-\frac{\theta}{1-\theta}}}{1-4\theta}, \end{align*} and so, $\beta_1, \ApproxVarExtrap{\beta}_2,\ApproxVarExtrap{\beta}_3$ are the coefficients appearing in Proposition \ref{Prop:ControlDecreasingStep} applied to this setup. We conclude as well that $\ApproxVarExtrap{R}_{F}^{i}$ is the function appearing in Proposition \ref{Prop:MainDecreasingStepPqCLT}. Similar developments for $H$ allow to extend the conclusion to $\ApproxVarExtrap{R}_{H}^{i,j}$. \item Finally, looking at the definition of $\varphi_F$ and $\Phi_F$ from Proposition \ref{Prop:MainDecreasingStepPqCLT} and \eqref{Eq:DerivativePhiHat} we get that \begin{multline*} \ApproxVarExtrap{\Phi}_F^{i,j}(x)=\frac{C_0^{-1}}{(\lambda-1)^2} \left( \lambda^{2} \int \langle \sigma^* D_y \phi_{F^i}, \sigma^* D_y \phi_{F^j} \rangle (x,y^1) \mu^x(dy^1) \right.\\ + \left.\int \langle \sigma^* D_y \phi_{F^i}, \sigma^* D_y \phi_{F^j} \rangle (x,y^2) \mu^x(dy^2) \right) ; \end{multline*} i.e. $\ApproxVarExtrap{\Phi}_F(x) = (\lambda^{2}+1)(\lambda-1)^{-2} \Phi_F(x)$. We get a similar result for $\ApproxVarExtrap{\Phi}_G$. We obtain the value $\ApproxVarExtrap{C}_\varphi$ given in the statement. The claim follows. \end{enumerate} \end{proof} \begin{MyRemark} $\ApproxVarExtrap{C}_\varphi$ is a constant multiplying the uncertainty coming from the decreasing step estimator. Since we would like this quantity as small as possible, having an explicit value for $\ApproxVarExtrap{C}_\varphi$ is very useful from a numerical point of view: we can choose $\lambda$ to minimize $\ApproxVarExtrap{C}_\varphi$. We get \[\lambda_* = 1+(\sqrt{3}+1)^{1/3}+(\sqrt{3}+1)^{-1/3}\approx 3.196\] inducing $\ApproxVarExtrap{C}_\varphi\approx 2.64 $. This is the initial additional cost that has to be paid for the extrapolation, making the EMsDS algorithm useful for large $n$, where the reduction in complexity of the EMsDS is enough to compensate for the higher error. \end{MyRemark} \section{Numerical results} \label{Sec:numresults} \subsection{Efficiency analysis} \label{Subsec:Efficiency} We can approximate the execution time of both algorithms, the original and extrapolated versions of the algorithm, by estimating the total number of operations needed to perform one path approximation of the effective equation \eqref{Eq:EffectiveEquation}. Note that since both algorithms share the same structure, a similar analysis is valid for both of them: the total cost $\kappa(n)$ of the algorithm with $n$ steps may be written as \[\kappa(n)= \left[\kappa_1(n,d_x,d_y)+\kappa_2(d_x)\right]n,\] where $\kappa_1$ stands for the cost coefficient estimation at each step of the decreasing Euler, and $\kappa_2$ for the cost of calculating the Euler iteration. The latter will be of order $O(d_x)$ in the ODE case and $O(d_x^2)$ for the SDE case. Let us focus now on $\kappa_1$. Both algorithms perform $M_1n^{1/(1-\theta)}$ iterations for approximating the diffusion $\DecrAvg{Y}$ and the calculation of estimators $\DecrAvg{F}, \DecrAvg{G}$. For the MsDS algorithm, each one of this iterations has a cost of $O(d_y d_x)$ in the $ODE$ case, or $O(d_y d_x^2)$ in the SDE case. In the latter, we need also to perform a Cholesky decomposition with a cost of $O(d_x^3)$ operations. Hence, \[ \kappa_1^{MsDS} (n,d_x,d_y ) = \begin{cases} O\left( d_yd_x n^{1/(1-\theta)} \right) &\text{ in ODE case } \\ O\left( [d_yd_x^2 + d_x^3] n^{1/(1-\theta)}\right) &\text{ in SDE case }\end{cases}.\] On the other hand, from the definition of the EMsDs algorithm, we get $\kappa_1^{EMsDS} \leq \lambda \kappa_1^{MsDS}$, and thus both share the same order of complexity, with the only difference that \emph{$\theta$ is allowed to be smaller in the extrapolated algorithm}. It may be more interesting to compare the \emph{efficiency} of both algorithms, that is the time spent to obtain a given error tolerance $\Delta$. We have from theorems \ref{Theo:ShortMainDecreasingStep} and \ref{Theo:ShortMainDecreasingStepInterp} that $\Delta(n):= O( n^{-1} ) $ for the ODE, and $\Delta(n):= O( n^{-1/2} )$ for the SDE case. Replacing the minimum possible $\theta$ values we obtain the complexity figures given in Table \ref{Tab:Efficiency}. \begin{table} \centering \caption{Minimal efficiency (operations for fixed error) of the basic and extrapolated algorithm for ODE and full SDE cases} \begin{tabular}{l|c|c|c|c} & ODE & ODE (extrapol.) & SDE & SDE (extrapol.)\\ \hline $\theta_{\min}$ & $1/2$ & $1/3$ & $1/3$ & $1/5$\\ $\tau_{\min}(\Delta)$ & $O(d_yd_x \Delta^{-3}) $ & $O(d_xd_y \Delta^{-2.5})$ & $O([d_x^2d_y+d_x^3] \Delta^{-5})$ & $O([d_x^2d_y+d_x^3] \Delta^{-4.5})$ \end{tabular} \label{Tab:Efficiency} \end{table} How do these figures compare with a straightforward Euler scheme applied to the original system? For the ODE case, an Euler scheme implemented for the original system \eqref{Eq:TheSystem} would require a total of $(dx+dy)\epsilon^{-1} \Delta^{-2}$ operations. Then the MsDS algorithm is more efficient if $\epsilon < \Delta(d_x\vee d_y)^{-1}$, and the EMsDS if $\epsilon < \Delta^{1/2} (d_x\vee d_y)^{-1}$. With respect to the algorithm presented in \cite{e_analysis_2005}, the efficiency is equivalent to the one obtained when using a weak scheme of order one for approximating the ergodic averages. The advantage of our method is that we have in addition to the rate of convergence an expression for a C.L.T. type result. In the SDE case, on the other hand, the proposed algorithm will be advantageous in the case in which $\epsilon < \Delta^3 (d_x\vee d_y)^{-1}$ for the MsDS version, and $\epsilon < \Delta^{2.5}(d_x\vee d_y)^{-1}$ for the EMsDS. In other words, our proposed algorithms will be more efficient in our regime of interest of a strong scale scale separation (i.e. when $\epsilon \rightarrow 0$). It should be remarked that the SDE case is not explicitly studied for the algorithm in \cite{e_analysis_2005}. \subsection{Numerical tests} \label{SubSec:NumericalTests} \subsubsection{A toy problem} Let us illustrate the main features of the algorithm by evaluating its behavior when used for solving a toy system for which we are able to obtain an exact solution. Consider \[dY^x_t = \left((|x|^2+1)^{-1/2} - Y^x_t \right) + \sqrt{2}d\tilde{W}_t, \] which is an Ornstein-Uhlenbeck system having a unique invariant measure distributed $\mathcal{N}\left( (|x|^2+1)^{-1/2}, 1 \right)$; and define the SDE system \[dX_t = F(X_t) dt + G(X_t) dW_t,\] with \begin{align*} f(x,y) & :=\begin{pmatrix} 1+y- (|x|^2+1)^{-1/2} \\ 1 \end{pmatrix}; & g(x,y) & := \sqrt{\frac{|x|^2+1}{2|x|^2+3} (y^2+1)} \begin{pmatrix} 1 & 0 \\ 1 & 1 \end{pmatrix}, \end{align*} with $F,G$ defined as before and where $\tilde{W}$ is a real Brownian motion independent of the planar Brownian motion $W$. The form of the assumed coefficients is chosen to satisfy the regularity and uniform bound hypothesis in $\ensuremath {(\mathcal{H}_{s.s.})}$ and $\ensuremath {(\mathcal{H}_{f.s.})}$ and to give a simple effective equation expression. In fact, it is easily verified that the exact effective equation is \[X_s = \begin{pmatrix} x_0^1 + s + W^1_s\\ x_0^2 + s + W^1_s+W^2_s\end{pmatrix} \] \begin{figure} \centering \includegraphics[width=0.44\textwidth] {./QQPlotSDEo1.pdf} \includegraphics[width=0.44\textwidth] {./QQPlotSDEiL1.pdf}\\ \includegraphics[width=0.44\textwidth] {./QQPlotSDEo2.pdf} \includegraphics[width=0.44\textwidth] {./QQPlotSDEiL2.pdf} \caption{QQ-plot comparing the rescaled errors in the simulation with $n=510$ and the theoretical limit distribution ( the reference line represents a perfect match). Left : SDE decreasing step. Right: SDE interpolated.} \label{fig:qqplot} \end{figure} \begin{figure} \centering \includegraphics[width=0.44\textwidth]{./StrongSquareErrorSteps.pdf} \includegraphics[width=0.44\textwidth]{./StrongSquareErrorTime.pdf} \caption{Left: $L_2$ error as a function of steps for the SDE case (log-log scale). Note that the estimated values for the slopes verify the rate of convergence for the algorithm in both implementations. Right: $L_2$ error as a function of execution time for the SDE case (log-log scale). Although a higher price must be payed for a small step number, the slope difference signals a change in the asymptotic order of convergence.} \label{fig:StrongErrorSteps} \end{figure} We will look at the numerical results of applying the decreasing step with sequence $\gamma_k = k^{-1/3}$ and the EMsDS version with sequence $\gamma_k = k^{-1/5}$ and $\lambda=3$. Let us examine the distribution of the error at a fixed time $T=1$ (i.e. $\zeta = \ApproxVar{X}_1 - X_1 $ ). Figure \ref{fig:qqplot} shows a Q-Q plot of the rescaled simulated errors $\sqrt{n} \zeta$ and the limit distribution error in the studied cases. As shown, the empirical distributions obtained after 1600 simulations with $n = 510$ verify the expected limit behavior. Figure \ref{fig:StrongErrorSteps} Left plots in a log-log scale the evolution of the $L_2$ error \[\zeta_{L_2} = \left(\EX{\sup\limits_{0\leq t \leq T} |\ApproxVar{X}_t - X_t |^2}\right)^{1/2}\] in function of the number of steps $n$, comparing both versions of the algorithm. The empirically obtained slope ( close to $-0.5$ in both cases ) represents the power of the approximation and is the one expected from the convergence theorems. We show as well in Figure \ref{fig:StrongErrorSteps} Right a comparison in the efficiency of both methods (measured as the error in terms of the execution time) of each one of the algorithms. The effect of the extrapolation in the cost of the algorithm is evidenced in the difference in slope of the empirical plot for both algorithms. Note that solving for $\Delta$ in Table \ref{Tab:Efficiency} we get $\Delta_{MsDS} = O(\tau^{-0.2})$ and $\Delta_{EMsDS} \approx O(\tau^{-0.222})$, values that are retrieved in the numerical experiment. It is worth observing the difference in the intercept of both lines, showing that the higher slope comes with a cost in the initial error. The conclusion drawn from the toy example may well be generalized: the user should consider implementing the extrapolated version only when requiring a very high precision on the approximation results. \subsubsection{Pricing in finance} We apply now the algorithm to a pricing problem in finance. Consider the mean-reverting corrected Heston's stochastic volatility model presented in \cite{fouque_fast_2011} and given by \begin{align*} & dX_t = rX_t dt + \Sigma_t X_t dW^{x}_t \\ & dY_t = \epsilon^{-1} Z_t (m-Y_t) dt + \nu \sqrt{2 Z_t \epsilon^{-1} } dW^y_t\\ & dZ_t = \kappa(\theta-Z_t)dt + \sigma \sqrt{Z_t} dW^z_t\\ & \Sigma_t = \sqrt{Z_t}(1+Y_t^2). \end{align*} where we assume $W_t^x,W^y_t,W^z_t$ are one-dimensional Brownian motions with correlations $\rho_{xy}$, $\rho_{xz}$ and $\rho_{yz}$. We suppose the model is already written in terms of the risk neutral probability measure with known parameters and initial conditions given by \begin{table}[h] \caption{Initial condition and parameters of the model. \label{Table:parameters}} \begin{tabular}{ | c c c | c c c c c c c c c| } \hline $x_0$ & $z_0$ & $y_0$ & $m$& $\nu$& $\kappa$ &$r$& $\theta$& $\sigma$ &$\rho_{xy}$ &$\rho_{yz}$ &$\rho_{xz}$ \\ \hline 100 &0.24 &0.06 & 0.06 & 1.0 & 1.0 &0.05 & 1.0 &0.39 & 0 & 0 & -0.33\\ \hline \end{tabular} \end{table} We are interested in pricing several types of options depending on the whole trajectory on this model. For this test, we price a floating strike Asian call (the payoff being $AC_{float}=S_T-T^{-1} \int S_t dt$ ) and a lookback call with floating strike (with payoff $LC_{float}=S_T-S_{\min}$). In this test, we compare the algorithm with a simple Euler scheme with different values for $\epsilon$. We carry out 6000 Montecarlo simulations. The results are presented in Table \ref{Table:simulation}. \begin{table}[h] \caption{Simulation values. \label{Table:simulation}} \begin{tabular}{ c c | c | c | c | c c c c c c| } \hline Method & $\epsilon$ & $n$ & $M(n)$ & $n \times M(n)$& Asian & Lookback \\ \hline\hline Euler & $10^{-3}$ & $5 \times 10^6$ & $1$ & $5 \times 10^4$ & 40.988 & 81.591\\ Euler & $10^{-3}$ & $ 10^7$ & $1$ & $1 \times 10^5$ & 40.503 & 81.256\\ Euler & $10^{-3}$ & $2\times 10^7$ & $1$ & $2 \times 10^5$ & 40.086 & 80.769\\ Euler & $10^{-4}$ & $5 \times 10^6$ & $1$ & $5 \times 10^5$ & 22.091 & 54.119\\ Euler & $10^{-4}$ & $ 10^7$ & $1$ & $1 \times 10^6$ & 21.897 & 53.806\\ Euler & $10^{-4}$ & $2\times 10^7$ & $1$ & $2 \times 10^6$ & 20.908 & 52.095\\ Euler & $10^{-5}$ & $5 \times 10^6$ & $1$ & $5 \times 10^6$ & 18.203 & 45.947\\ Euler & $10^{-5}$ & $ 10^7$ & $1$ & $1 \times 10^7$ & 15.164 & 39.123\\ Euler & $10^{-5}$ & $2\times 10^7$ & $1$ & $2 \times 10^7$ & 20.659 & 51.240\\ \hline MsDS & - & 50 & 3540 & $1.77 \times 10^5$ & 20.738 & 47.920\\ MsDS & - & 100 & 10010 & $1 \times 10^6$ & 20.681 & 48.841\\ MsDS & - & 200 & 28290 & $5.66 \times 10^6$ & 20.669 & 49.557\\ \hline \end{tabular} \end{table} Note that the system does not satisfy all the hypothesis $\ensuremath {(\mathcal{H}_{f.s.})}$ and $\ensuremath {(\mathcal{H}_{s.s.})}$, particularly it fails to satisfy the boundedness of the coefficients with respect to the slow variables, and the uniform ellipticity hypothesis. Nevertheless, the MsDS algorithm seems to work even under these relaxed conditions, and, in addition, appears to be more stable than the algorithm using small values of $\epsilon$. Note as well that for similar values of total operations (represented by the column $n\times M(n)$), the MsDS algorithm gives better results. \section*{Acknowledgements} The author would like to thank Fran\c{c}ois Delarue for his help and support during the preparation of this work, and the anonymous referee for his suggestions that greatly improved the paper.
\section{Introduction} The need to cure the infinities plaguing the quantum fields was the first motivation to enlarge the Lorentz symmetry to include the noncommutative algebra [1, 2]. However, in recent decade interests in noncommutative theories gained a considerable attention because of the discoveries in string theory which imply that noncommutative space-time may be an inherent part of the high energy (Planck scale) physics. A space is noncommutative if it's coordinates satisfy \begin{equation}\label{1} [\widehat{x}^\alpha,\widehat{x}^\beta]=i\theta^{\alpha\beta},\quad\quad\theta^{\alpha\beta}=-\theta^{\beta\alpha}. \end{equation} As a result of "$\theta$-deformation" of the algebra of space-time coordinates one must replace the usual product among the fields with Weyl-Moyal product or $\star$-product [2], i.e. \begin{equation}\label{1} \phi_1(\widehat x)\star\phi_2(\widehat x)\equiv \lim_{x\rightarrow y}e^{\frac{i}{2}\theta^{\mu\nu}\partial^x_\mu\partial^y_\nu}\phi_1(x)\phi_2(y). \end{equation} From the Feynman's rules point of view, the only effect of the $\star$-product is to modify the $n$-point interaction vertices (3$\leq n$) by the phase factor [2] \begin{equation}\label{1} \tau(p_1,\ldots,p_n)=e^{-\frac{i}{2}\sum^n_{a<b}\,p_a\wedge p_b}. \end{equation} where $p_a\wedge p_b=\theta_{\mu\nu}p^\mu_a p^\nu_b$. Here the momentum flow of the $a$-th field into the vertex is denoted by $p_a$. In the case of gravitational field the effect of noncommutating coordinates is to modify the Newton potential as [3-6] \begin{equation}\label{1} V_\theta=-G\frac{m_am_b}{r}-G\frac{m_am_b}{2r^3}\,\textbf{L}\cdot\boldsymbol{\theta}+O(\theta^2). \end{equation} where $\textbf{L}$ is the particle's angular momentum. The possible effects of the $\theta$-deformed gravitational potential (4) on the celestial dynamics is considered by several authors [3-5].\\\ In present work we shall consider the gravitational radiation of a two body system in a circular motion with deformed potential between the massive bodies. We follow a quantum field theoretic approach to derive the potential (4) in next section. In section 3, we will consider the classical gravitation two body problem and shall re-derive the deformed potential using the so called Bopp shift. In section 4, the gravitational radiation power and the period decrease rate are calculated for the system up to first order in noncommutativty parameter $\theta$. Finally, in section 5, the period decay of the model is compared with the observational data of binary pulsar PSR 1913+16 to obtain a bound on the noncommutativty parameter. \section{Deformed Newton Potential} In a NC flat background the interaction between gravitational and scalar fields is [6, 7] \begin{equation}\label{1} \mathcal{L}_{int}=-\frac{\gamma}{2}{h}^{\mu\nu}\star \Big[\partial_\mu{\phi}\star\partial_\nu{\phi}-\frac{1}{2}\big(\partial_\alpha{\phi}\star\partial^\alpha{\phi}-m^2\phi\star\phi\big)\Big],\quad \gamma^2=32\pi G. \end{equation} Therefore, with the aid of formula (3) we obtain the deformed momentum space 2 scalar-1 graviton vertex factor as \begin{equation}\label{1} \tau_{\alpha\beta}^\theta(p,p')=-\frac{i\gamma}{2}(p_\alpha p'_\beta+p'_\alpha p_\beta-\eta_{\alpha\beta}p\cdot p')e^{\frac{i}{2}\textbf{p}\wedge \textbf{p}'}. \end{equation} where we have assumed $\theta_{\mu 0}=0$ implying $\theta^{ij}p^i q^j\rightarrow\textbf{p}\wedge \textbf{q}$ to avoid the problematic features of the noncommutative models [2]. Now, let us look at a typical two-body scattering mediated by a graviton. For the spinless particles with masses $m_a$ and $m_b$ the scattering amplitude is [6] {\setlength\arraycolsep{2pt} \begin{eqnarray}\label{8} \mathcal M_\theta&=&\tau^{\mu\nu}_\theta(p_a,p'_a)D_{\mu\nu,\alpha\beta}(p_a-p'_a)\tau^{\alpha\beta}_\theta(p_b,p'_b),\\\nonumber &=&\frac{4\pi G}{(p_a-p'_a)^2}\bigg(\Big[(p_a+p_b)^2-m^2_a-m^2_b\Big]^2+\Big[(p_a-p'_b)^2-m^2_a-m^2_b\Big]^2\\\nonumber &-&\Big[(p'_a-p_a)^2+4m^2_am^2_b\Big]^2\bigg)e^{i \textbf{p}\wedge (\textbf{p}- \textbf{p}')}, \end{eqnarray}} where the momentum-space graviton propagator is \begin{equation}\label{1} D_{\mu\nu\alpha\beta}(q)=-\frac{i}{2q^2}(\eta_{\mu\alpha}\eta_{\nu\beta}+\eta_{\mu\beta}\eta_{\nu\alpha}-\eta_{\mu\nu}\eta_{\alpha\beta}). \end{equation} In the non-relativistic limit we have {\setlength\arraycolsep{2pt} \begin{eqnarray}\label{8} (p'_a-p_a)^2&\approx& - \textbf{q}^{\,2},\\ (p_a+p_b)^2&\approx&(m_a+m_b)^2,\\ (p_a-p'_b)^2&\approx&(m_a-m_b)^2+\textbf{q}^{\,2}. \end{eqnarray}} By substituting (9)-(11) in (7) we find the deformed gravitational potential {\setlength\arraycolsep{2pt} \begin{eqnarray}\label{8} U_{\theta}(\textbf{x})&=&-\frac{1}{4m_am_b}\int\frac{d^3q}{(2\pi)^3}\mathcal M_{\theta}(\textbf{q})e^{i\textbf{q}\cdot \textbf{x}},\\\nonumber &=&-G\frac{m_am_b}{\sqrt {({x^i-\frac{1}{2}\theta^{ij}p^j})({x^i-\frac{1}{2}\theta^{ik}p^k })}},\\\nonumber &=&-\frac{\kappa}{r}-\frac{\kappa}{2r^3}\,\textbf{L}\cdot\boldsymbol{\theta}+O(\theta^2). \end{eqnarray}} with $\kappa=Gm_am_b$. \section{Two body Problem in NC Gravity} The hamiltonian describing two particles $a$ and $b$ interacting via the Newton potential is \begin{equation}\label{1} H=\frac{\textbf{p}^2_a}{2m_a}+\frac{\textbf{p}^2_b}{2m_b}-\frac{\kappa}{|\widehat{\textbf{x}}_a-\widehat{\textbf{x}}_b|}. \end{equation} The classical canonical structure of the above system in NC space has the form {\setlength\arraycolsep{2pt} \begin{eqnarray} \{\widehat{x}_a^i,\widehat{x}_a^j\}&=&\{\widehat{x}_b^i,\widehat{x}_b^j\}=\theta^{ij},\\\ \{\widehat{x}_a^i,\widehat{p}_a^j\}&=&\{\widehat{x}_b^i,\widehat{p}_b^j\}=\delta^{ij}. \end{eqnarray}} where we have used the correspondence $\frac{1}{i}[A,B]\rightarrow\{A,B\}$ to achieve the classical canonical structure from its quantum counterpart [3, 4]. We introduce the new set of coordinates {\setlength\arraycolsep{2pt} \begin{eqnarray} \textbf{X}&=&\textbf{x}_{a}-\textbf{x}_b,\\\ \textbf{X}_c&=&\frac{m_a\textbf{x}_{a}+m_b\textbf{x}_b}{m_a+m_b}, \end{eqnarray}} to rewrite (13) as \begin{equation}\label{1} H=\frac{\textbf{p}^2_c}{2(m_a+m_b)}+\frac{\textbf{p}^2_{X}}{2\mu}-\frac{\kappa}{|\widehat{\textbf{X}}|}, \end{equation} with classical canonical structure given by {\setlength\arraycolsep{2pt} \begin{eqnarray} [\widehat{X}^i,\widehat{X}^j]&=&2\theta^{ij},\\\ [\widehat{X}^i,\widehat{P}^j_X]&=&\delta^{ij}. \end{eqnarray}} Now, the so-called Bopp shift, i.e. \begin{equation}\label{1} \widehat{X}^i\rightarrow X^i=\widehat{X}^i+\theta^{ij}P^j_X, \end{equation} allows one to introduce the variable which fulfills the standard canonical structure {\setlength\arraycolsep{2pt} \begin{eqnarray} [{X}^i,{X}^j]&=&0,\\\ [{X}^i,{P}^j_X]&=&\delta^{ij}. \end{eqnarray}} By assuming that the center of mass is fixed, i.e. $\textbf{p}_c=0$, and on substituting $\theta\rightarrow 2\theta$ (c.f. (19)) the deformed Hamiltonian becomes \begin{equation}\label{1} H_\theta=\frac{\textbf{p}^2_{X}}{2\mu}-\frac{\kappa}{R}-\frac{\kappa}{R^3}\,\textbf{L}\cdot\boldsymbol{\theta},\quad {R=|\textbf{X}}|. \end{equation} For a circular motion i.e. $\dot R=0$ the equation of motion yields \begin{equation}\label{1} \frac{\partial H_\theta}{\partial R}=\mu R^3{\omega^2_\theta}-\kappa{\mu\theta\omega_\theta\cos\alpha}-{\kappa}=0. \end{equation} where we have used $L=\mu R^2\omega_0$ with $\omega^2_0=G\frac{m_a+m_b}{R^3}=\frac{\kappa}{\mu R^3}$. Here $\alpha$ denotes the angle between $\boldsymbol{\theta}$ and $\textbf{L}$. From (25) one finds the angular velocity of the system as {\setlength\arraycolsep{2pt} \begin{eqnarray} {\omega_\theta}&=&\frac{1}{2}\omega^2_0{\mu\theta\cos\alpha}+\sqrt{\omega^2_0+\frac{1}{4}\omega^4_0{\mu^2\theta^2\cos^2\alpha}},\\\nonumber &=&\omega_0+\frac{1}{2}\omega^2_0{\mu\theta\cos\alpha}+O(\theta^2). \end{eqnarray}} \section{Period Decay in a Compact Binary} For a binary system located at $X^3=0$ plane, the coordinates of the bodies $a$ and $b$ circulating around the center of mass, are {\setlength\arraycolsep{2pt} \begin{eqnarray} X^1_a &=&- X^1_b =\mu R\cos\omega_\theta t ,\\ X^2_a &=& -X^2_b=\mu R\sin\omega_\theta t , \\ X^3_a &=& X^3_b =0. \end{eqnarray}} The total gravitational power radiated by the system is $P=\frac{G}{45}\langle\dddot{D}^{ij} \dddot{D}^{ij} \rangle$ where the quadruple moment is [8] {\setlength\arraycolsep{2pt} \begin{eqnarray}\label{1} D^{11}&=&\mu R^2(3\cos^2\omega_\theta t -1),\\\ D^{22}&=&\mu R^2(3\sin^2\omega_\theta t -1),\\\ D^{21}&=&D^{12}=3\mu R^2\sin\omega_\theta t\cos\omega_\theta t, \\\ D^{33}&=&-\mu R^2. \end{eqnarray}} The radiated power by the both particles, $P_\theta=P_{\theta a}+P_{\theta b}$, becomes {\setlength\arraycolsep{2pt} \begin{eqnarray}\label{1} P_\theta=-\frac{dE_\theta}{dt}&=&\frac{32}{5c^5}G\mu^2R^4\omega^6_\theta,\\\nonumber &\simeq&\frac{32}{5c^5}G\mu^2R^4\omega^6_0+\frac{96}{5c^5}G\mu^3R^4\omega^7_0\theta\cos\alpha. \end{eqnarray}} The energy of system is {\setlength\arraycolsep{2pt} \begin{eqnarray} E_\theta&=&\frac{1}{2}\mu\omega_\theta^2R^2-\frac{\kappa}{R}-\frac{\kappa}{R}\mu\omega_0\theta\cos\alpha,\\\nonumber &\simeq&\frac{1}{2}\mu\omega_0^2R^2-\frac{\kappa}{R}+\frac{1}{2}\mu^2\omega_0^3R^2\theta\cos\alpha-\frac{\kappa}{R}\mu\omega_0\theta\cos\alpha,\\\nonumber &=&-\frac{\kappa}{2R}-\frac{1}{2}\sqrt{\frac{\mu\kappa^3}{R^5}}\theta\cos\alpha. \end{eqnarray}} Therefore the rate of energy lose takes the form \begin{equation}\label{1} -\frac{dE_\theta}{dt}=-\frac{\kappa}{2R^2}\Big(1+\frac{5}{2}\omega_0\theta\mu\cos\alpha\Big)\frac{dR}{dt}. \end{equation} Thus by equating the left hand side of (36) to (34) one obtains \begin{eqnarray}\label{1} \frac{dR}{dt}=-\frac{64}{5c^5}\Big(\frac{G}{R}\Big)^3{m_am_b(m_a+m_b)}\,\frac{1+3\omega_0{\mu\theta\cos\alpha}}{1+\frac{5}{2}\omega_0\mu\theta\cos\alpha}. \end{eqnarray} For $\theta=0$ the above expression coincides with the well-known textbook result [8]. From (37) and by virtue of $\tau\equiv\dot T=3\pi\sqrt{\frac{\mu R}{\kappa}}\dot R$ we obtain the rate of period decay as {\setlength\arraycolsep{2pt} \begin{eqnarray}\label{1} \tau_\theta&=&-\frac{192}{5}\frac{\pi m_am_b}{(m_a+m_b)^{\frac{1}{3}}}\Big(\frac{2\pi G}{T}\Big)^{\frac{5}{3}}\frac{1+3\omega_0{\mu\theta\cos\alpha}}{1+\frac{5}{2}\omega_0\mu\theta\cos\alpha},\\\nonumber &\simeq&\tau_0+\Delta\tau. \end{eqnarray}} where $\Delta\tau=\frac{1}{2}\tau_0\omega_0\mu\theta\cos\alpha$. Again, for $\theta=0$ we are left with the standard result for the rate of (circular) orbit decay [9]. \section{PSR 1913+16 Binary System} The masses of the pulsar and its companion in PSR 1913+16 binary system are {\setlength\arraycolsep{2pt} \begin{eqnarray} m_p &=& 1.44 \,M_\odot,\\ m_c &=& 1.38 \,M_\odot. \end{eqnarray}} and the eccentricity of system is $e=0.61$. For non-circular orbit, i.e. for $e\neq 0$ case, the orbit decay rate $\tau_0$ includes the factor $f(e)$, which satisfies $f(0.61)=11.85$ [9]. The theoretical and observed values for the orbit decay rates, the reduced mass and period of the system, respectively are [9] {\setlength\arraycolsep{2pt} \begin{eqnarray}\label{1} \tau_0&=&-2.42\times 10^{-12}\,\textrm{sec/sec},\\\ \tau_0^{obs}&=&-2.40 \times 10^{-12}\, \textrm{sec/sec},\\\ \mu&=&0.7\,M_\odot=1.39\times 10^{30}\, \textrm{kg},\\\ T&=&27898.56\ \textrm{sec}. \end{eqnarray}} So, by assuming $\cos\alpha=1$, from the constraint \begin{equation}\label{1} \Delta \tau < \big|\tau_0-\tau_0^{obs}\big|, \end{equation} we find \begin{equation}\label{1} \theta <1.6323\times10^{-29}. \end{equation} One must note that the above result for the noncommutativity parameter is valid as an estimation since the left hand side of constraint (45) does not include the factor accounting for the eccentricity of the system due to the fact that our analysis is restricted to the circular orbit.
\section{introduction} \subsection{Background and motivation} Let $X$ be a simply laced Dynkin diagram with the index set $I=\{1,2,\cdots, r\}$. For a family of variables $\{ Q^{(a)}_{m}|a\in I, m\in \mathbb{Z}_{\geq 0} \}$, consider recurrences given by \begin{equation}\label{4:Qsys} \left(Q^{(a)}_{m}\right)^2=\prod _{b\in I} \left(Q^{(b)}_{m}\right)^{\mathcal{I}(X)_{ab}}+Q^{(a)}_{m-1}Q^{(a)}_{m+1} \quad, m\ge 1 \end{equation} where $\mathcal{I}(X)$ denotes the adjacency matrix of $X$. We call it the unrestricted $Q$-system of type $X$. Throughout the paper, we will use the boundary conditions $Q^{(a)}_{0} =1$ for all $a\in I$. Let $k\geq 1$ be an integer. We are interested in finding complex solutions of the $Q$-system satisfying another set of boundary conditions $Q^{(a)}_{k}=1$ for all $a\in I$. We define the level $k$ restricted $Q$-system of type $X$ to be the system of equations \begin{equation} \left\{ \begin{array}{lll} Q^{(a)}_{0} =1 & a\in I \\ \left(Q^{(a)}_{m}\right)^2=\prod _{b\in I} \left(Q^{(b)}_{m}\right)^{\mathcal{I}(X)_{ab}}+Q^{(a)}_{m-1}Q^{(a)}_{m+1} & 1\le m <k, a\in I\\ Q^{(a)}_{k} =1 & a\in I \end{array} \right. \end{equation} in variables $\{Q^{(a)}_{m}| a\in I , 0\leq m \leq k\}$. One reason to consider it comes from Nahm's conjecture about modularity of $q$-hypergeometric series. See \cite{Nahm, Keegan:2007zq, 2009arXiv0905.3776N} for instance. Another reason to study it can be found in dilogarithm identities for conformal field theories \cite{springerlink:10.1007/BF01840426, MR2804544}. To motivate our investigation, we give a brief exposition of dilogarithm identities related to our main results. It is known that there exists a special unique solution of the level $k$ restricted $Q$-system possessing positivity and some additional properties as follows : \begin{thm}\label{5:Quni} Let $X$ be a Dynkin diagram of type $ADE$ of rank $r$. There exists a unique solution $\mathbf{z}=(z^{(a)}_{m})$ of the level $k$ restricted $Q$-system of type $X$ satisfying $z^{(a)}_{m}>0$ for $0\leq m \leq k$ and $a\in I$. For all $a\in I$, the following properties hold : \begin{enumerate} \item (symmetry) $z^{(a)}_{m}=z^{(a)}_{k-m}$ for $0\leq m \leq k$, \item (unimodality) $z^{(a)}_{m-1}<z^{(a)}_{m}$ for $1\le m \le \lfloor\frac{k}{2}\rfloor$ where $\lfloor x\rfloor$ is the floor function. \end{enumerate} \end{thm} See \cite[Theorem 5.3.6]{chlee2012} for a proof. We call the solution $\mathbf{z}=(z^{(a)}_{m})$ characterized in Theorem \ref{5:Quni} the positive solution of the level $k$ restricted $Q$-system. The Rogers dilogarithm function is defined by $$L(x)=-\frac{1}{2}\int_{0}^{x}\frac{\log(1-y)}{y}+\frac{\log(y)}{1-y}\,dy$$ for $x\in (0,1)$. We set $L(0)=0$ and $L(1)=\pi^2/6$ so that $L$ is continuous on $[0,1]$. \begin{thm}\cite{MR2804544} Let $X$ be a simply laced Dynkin diagram of rank $r$ and $\mathfrak{g}$ the corresponding simple Lie algebra. For $1\leq m \leq k-1$ and $a\in I$, let $$x_{m}^{(a)}=\frac{\prod_{b\in I} (z_{m}^{(b)})^{\mathcal{I}(X)_{ab}}}{(z_{m}^{(a)})^2}$$ where $\mathbf{z}=(z^{(a)}_{m})$ is the positive solution of the level $k$ restricted $Q$-system of type $X$. The following dilogarithm identity holds : \begin{equation}{\label{dilogcft}} \frac{6}{\pi^2}\sum_{a \in I}\sum_{m=1}^{k-1}L(x_{m}^{(a)})=\frac{k \dim \mathfrak{g}}{h+k}-\operatorname{rank} \mathfrak{g}=\frac{(k-1)h r}{h+k} \end{equation} where $h$ denotes the Coxeter number of $\mathfrak{g}$. \end{thm} For a physical interpretation of the rational number on the right hand side of (\ref{dilogcft}), see \cite[Theorem 5.2]{1751-8121-44-10-103001} and references given there. In this paper, we will study the positive solution of the level $k$ restricted $Q$-system of type $X$ using Lie theory. Before stating our main results, let us set up notation and terminology. \subsection{Notation} Let $X$ be a simply laced Dynkin diagram and $\mathcal{C}=(a_{ij})$ the Cartan matrix. Let $\mathfrak{g}$ be the corresponding simple Lie algebra of rank $r$ and $\mathfrak{h}$ its Cartan subalgebra. We denote the dual space of $\mathfrak{h}$ by $\mathfrak{h}^{*}$ and use the symbol $\langle \cdot ,\cdot \rangle$ to denote the natural pairing between $\mathfrak{h}$ and $\mathfrak{h}^{*}$. Let $\Phi\subset \mathfrak{h}^{*}$ be the root system. We denote the set of positive roots by $\Phi^{+}$ and the set of simple roots by $\Pi=\{\alpha_{i}|i\in I\}$. We will write $\alpha>0$ if $\alpha\in \Phi^{+}$ and $\alpha\geq \beta$ if $\alpha-\beta \in \Phi^{+}$ or $\alpha=\beta$. For $\alpha=\sum_{i=1}^{r}c_i\alpha_{i}\in \Phi$, we define its height, denoted by $\operatorname{ht} \alpha$, to be $\sum_{i=1}^{r} c_i$. Let $\theta=\sum_{i=1}^{r}a_i\alpha_{i}$ be the highest root. We call $a_i$ the marks and set $a_0=1$. See Figures \ref{pic:D} and \ref{pic:extD} for $X=D_r$. We denote the Coxeter number by $h$, which is given by $1+\operatorname{ht} \theta=\sum_{i=0}^{r}a_i$. Let $(\cdot |\cdot )$ be the standard symmetric bilinear form on $\mathfrak{h^{*}}$ normalized by requiring that $(\theta|\theta)=2$. Let $Q$ be the root lattice and $P$ be the weight lattice. The coroot lattice, denoted by $Q^{\vee}$ is the $\mathbb{Z}$-dual of the weight lattice $P$. We will choose the basis $\Pi^{\vee}=\{h_{i}\in \mathfrak{h}|i\in I\}$ of the coroot lattice so that $\langle \alpha_{i},h_j\rangle=a_{ji}$. Let $\{\omega_{i}\in P|i\in I\}$ be the dual basis of $P$ for $\Pi^{\vee}$ so that $\langle\omega_{i},h_{j}\rangle=\delta_{ij}$. We call $\omega_{i}$ the fundamental weights. A dominant weight is an element of $P_{+}=\{\sum_{i=1}^{r}\lambda_{i}\omega_{i}\in P|\lambda_{i}\ge 0, i\in I\}$. We call $\rho=\sum_{i=1}^{r}\omega_{i}\in P$ the Weyl vector. We have the group algebra $\mathbb{C}[P]$ with $\mathbb{C}$-basis of elements of the form $e^{\lambda}$, $\lambda \in P$. We can regard $e^{\lambda}$ as a function defined on $\mathfrak{h}^{*}$ by $\mu \mapsto e^{2\pi i (\lambda|\mu)}$. For a dominant weight $\lambda\in P_{+}$, the character $\chi_{\lambda}\in \mathbb{C}[P]$ of an irreducible representation $V$ of highest weight $\lambda$ is defined to be $$\sum_{\lambda' \in \mathfrak{h}^{*}} (\dim{V_{\lambda'}})e^{\lambda'}$$ where $V_{\lambda'}$ denotes the weight space corresponding to $\lambda' \in \mathfrak{h}^{*}$. We will regard $\chi_{\lambda}$ as a function on $\mathfrak{h}^{*}$. \begin{figure}\label{pic:D} \begin{tikzpicture} \begin{scope}[start chain] \dnode{1} \dnode{2} \dnode{3} \dydots \dnode{r-2} \dnode{r-1} \end{scope} \begin{scope}[start chain=br going above] \chainin(chain-2); \dnodebr{0}; \end{scope} \begin{scope}[start chain=br2 going above] \chainin(chain-5); \dnodebr{r}; \end{scope} \end{tikzpicture} \caption{The extended Dynkin diagram of type $D_r^{(1)}$ and the labeling of the nodes} \end{figure} \subsection{Statement of the KNS conjecture and the main theorem} Let $q$ be a non-zero complex number which is not a root of unity. The Kirillov-Reshetikhin (KR) modules form a special class of finite dimensional modules of the quantum affine algebra $U_{q}(\hat{\mathfrak{g}})$ and they are parametrized by $a\in I$, $m\in \mathbb{Z}_{\geq 0}$ and $u\in \mathbb{C}$. Since the quantized universal enveloping algebra $U_{q}(\mathfrak{g})$ is contained in $U_{q}(\hat{\mathfrak{g}})$ as a subalgebra, for a given KR module $W^{(a)}_{m}(u)$, we can get the finite dimensional $U_{q}(\mathfrak{g})$-module ${\rm res}\, W^{(a)}_{m}(u)$ by restriction. The important point to note here is the fact that the classical characters $Q^{(a)}_m$ of ${\rm res}\, W^{(a)}_{m}(u)$ for $a\in I$ and $m\in \mathbb{Z}_{\geq 0}$ satisfy the unrestricted $Q$-system, which was first stated in \cite{Kirillov1990} and later proved in \cite{MR1993360} and \cite{MR2254805}. The character $Q^{(a)}_m$ can be expanded into a sum of characters of irreducible modules of $\mathfrak{g}$ as \begin{equation}\label{4:Qdecomposition2} Q^{(a)}_{m}=\sum_{\omega \in P_{+}}Z(a, m,\omega)\chi_{\omega} \end{equation} where $Z(a, m,\omega)$ is a certain non-negative integer with $Z(a, m,m\omega_a)=1$. For example, when $X=A_r$, we have $Q^{(a)}_{m}=\chi_{m \omega_{a}}$ for $a\in I$ and $m\in \mathbb{Z}_{\geq 0}$ and they satisfy the unrestricted $Q$-system of type $A_r$. For $X=D_r$, it is given by \begin{equation}\label{6:qdecompD} Q^{(a)}_{m}= \left\{ \begin{array}{lll} \sum \chi_{k_{a}\omega_{a}+k_{a-2}\omega_{a-2}+\cdots + k_{1} \omega_{1}}& 1\leq a \leq r-2, a\equiv 1 \pmod 2, \\ \sum \chi_{k_{a}\omega_{a}+k_{a-2}\omega_{a-2}+\cdots + k_{0} \omega_{0}} & 1\leq a \leq r-2, a\equiv 0 \pmod 2, \\ \chi_{m \omega_{a}} & a=r-1,r \end{array} \right. \end{equation} where $\omega_0=0$ and the summation is over all nonnegative integers satisfying $k_{a}+k_{a-2}+\cdots + k_{1}=m$ for $a$ odd and $k_{a}+k_{a-2}+\cdots + k_{0}=m$ for $a$ even. For a treatment of more general cases, see \cite[Appendix A]{MR1745263}, \cite[Section 13]{1751-8121-44-10-103001} and references given there. If we regard $Q^{(a)}_{m}$ as a sum of characters given by (\ref{4:Qdecomposition2}), we can specialize it at the element $\frac{\rho}{h+k}\in \mathfrak{h}^{*}$. For each $a\in I$ and $m\in \mathbb{Z}_{\geq 0}$, we define $z^{(a)}_{m}$ by $$z^{(a)}_{m}=Q^{(a)}_{m}(\frac{\rho}{h+k})=\sum_{\omega \in P_{+}}Z(a, m,\omega)\chi_{\omega}(\frac{\rho}{h+k}).$$ This yields a solution of the unrestricted $Q$-system and we call it the quantum dimension solution of the $Q$-system. In \cite{1751-8121-44-10-103001}, it has been conjectured that it gives the positive solution of the level $k$ restricted $Q$-system and satisfies some additional level truncation properties. \begin{conjecture} \cite[Conjecture 14.2.]{1751-8121-44-10-103001}\label{KNSconj} Let $z^{(a)}_{m}=Q^{(a)}_{m}(\frac{\rho}{h+k})$ for $a\in I$ and $m\in \mathbb{Z}_{\geq 0}$. For all $a\in I$, the following properties hold : \begin{enumerate} \item (positivity) $z^{(a)}_{m}>0$ for $0\leq m \leq k$, \item (symmetry) $z^{(a)}_{m}=z_{k-m}^{(a)}$ for $1\leq m \leq k-1$, \item (unit boundary condition) $z^{(a)}_{k}=1$, \item (unimodality) $z_{m-1}^{(a)}<z^{(a)}_{m}$ holds true for $1\le m \le \lfloor\frac{k}{2}\rfloor$ where $\lfloor x\rfloor$ is the floor function, \item (occurrence of 0) $z^{(a)}_{k+1}=z^{(a)}_{k+2}=\cdots =z^{(a)}_{k+h-1}=0$. \end{enumerate} \end{conjecture} We call Conjecture \ref{KNSconj} the KNS conjecture. The conjecture was originated from some unproven claims in \cite{springerlink:10.1007/BF01840426} whose motivation can be found in the study of thermodynamic properties of the RSOS models \cite{MR1017122}. Then it was subsequently formulated as above in \cite{MR1304818} with more general specializations. The conjecture had been proved only in the case of type $A_r$. In this paper, we will prove the following. \begin{thmnn}\label{mainthmDr} The KNS conjecture is true for $X=D_r$. Moreover, $z^{(a)}_{k+h}=1$ for $1\le a\le r-2$, $z^{(r-1)}_{k+h}=z^{(r)}_{k+h}=1$ when $r\equiv 0,1 \pmod 4$ and $z^{(r-1)}_{k+h}=z^{(r)}_{k+h}=-1$ when $r\equiv 2,3 \pmod 4$. \end{thmnn} Since our proof crucially depends on (\ref{4:Qdecomposition2}), we cannot properly deal with the exceptional types where the decompositions are still largely conjectural. There are more general $Q$-systems including non-simply laced types, for which we need to modify (\ref{4:Qsys}) into a slightly more complicated form. Thus we focus on the case of type $D_r$ to see the central idea clearly. A proof for other classical types will be given in a forthcoming paper. This result will be divided into several parts and will be proved in Theorem \ref{main1}, \ref{main2} and \ref{main3}. The most tricky part lies in proving the unit boundary condition. Once we prove it, many results follow from Theorem \ref{5:Quni}. The key ingredients of our proof are the affine Weyl group symmetry and the extended Dynkin diagram symmetry of quantum dimensions of the affine weights obtained by suitable affinizations of classical weights. Although these are well-known concepts, they have not been effectively employed to attack our problem. In Section \ref{sec:qdim} we review necessary results about quantum dimensions. Section \ref{sec:comp} contains some preliminary calculations involving the affine Weyl group, which will be used in Section \ref{sec:mainpf} where we prove our main results. \section{Review on quantum dimensions}\label{sec:qdim} \begin{figure}\label{pic:extD} \begin{tikzpicture} \begin{scope}[start chain] \dnode{1} \dnode{2} \dnode{2} \dydots \dnode{2} \dnode{1} \end{scope} \begin{scope}[start chain=br going above] \chainin(chain-2); \dnodebr{1}; \end{scope} \begin{scope}[start chain=br2 going above] \chainin(chain-5); \dnodebr{1}; \end{scope} \end{tikzpicture} \caption{The extended Dynkin diagram of type $D_r^{(1)}$ and the marks.} \end{figure} In this section we summarize some of the standard facts on quantum dimensions without proofs. For a thorough treatment we refer the reader to \cite[Section 16.3]{philippe1997conformal}. \begin{definition} Let $\hat{P}$ be the lattice generated by $\hat{\omega}_0,\hat{\omega}_1,\cdots, \hat{\omega}_r$. We define $\hat{P}^k$ to be $\{\sum_{i=0}^{r}\lambda_{i}\hat{\omega}_{i}\in \hat{P}|\sum_{i=0}^{r}a_{i}\lambda_{i}=k\}$ where $a_0=1$. We will denote the set $\{\sum_{i=0}^{r}\lambda_{i}\hat{\omega}_{i}\in \hat{P}^k|\lambda_{i}\ge 0\}$ by $\hat{P}_{+}^{k}$. For a weight $\omega=\sum_{i=1}^{r}\lambda_{i}\omega_{i}\in P$, we define its level $k$ affinization $\hat{\omega}\in \hat{P}^k$ to be $\sum_{i=0}^{r}\lambda_{i}\hat{\omega}_{i}$. Note that $\lambda_0\in \mathbb{Z}$ is uniquely determined by the requirement $\hat{\omega} \in \hat{P}^k$. Let $\alpha_0=-\theta$ and $\hat{\alpha}_j=\sum_{i=0}^{r} (\alpha_j|\alpha_i) \hat{\omega}_{i}$. We define the fundamental reflections $s_0,s_1,\cdots s_r$ on $\hat{P}$ linearly by $$s_i \hat{\omega}_j=\hat{\omega}_j -\delta_{ij}\hat{\alpha}_i$$ where $\delta_{ij}$ denotes the Kronecker delta. They generate the affine Weyl group $W$. The signature of $w\in W$ will be denoted by $(-1)^{\ell(w)}$ where $\ell(w)$ is the length of $w\in W$. Let $\hat{\rho}=\sum_{i=0}^{r}\hat{\omega}_{i}\in \hat{P}$. We define the shifted affine Weyl group action on the set $\hat{P}$ by $$w\cdot \hat{\lambda}=w(\hat{\lambda}+\hat{\rho})-\hat{\rho}$$ for $w\in W$. \end{definition} \begin{definition} Let $\lambda \in P$ and $\hat{\lambda}\in \hat{P}^{k}$ be its level $k$ affinization. The quantum dimension \index{quantum dimension} or $q$-dimension \index{$q$-dimension} of $\hat{\lambda}$ is defined by \begin{equation}\label{6:qdimdef} \mathcal{D}_{\hat{\lambda}}=\chi_{\lambda}\left(\frac{\rho}{h+k}\right)= \frac{\prod_{\alpha>0}\sin \frac{\pi(\lambda+\rho|\alpha)}{h+k}}{ \prod_{\alpha>0}\sin \frac{\pi (\rho|\alpha)}{h+k}}. \end{equation} \end{definition} \begin{theorem} \label{6:qdimpos} Let $\lambda=\sum_{i=1}^{l} \lambda_i \omega_i\in P_{+}$ be a dominant weight such that $\sum_{i=1}^{l}a_i\lambda_i \leq k$. For its level $k$ affinization $\hat{\lambda} \in \hat{P}_{+}^{k}$, $\mathcal{D}_{\hat{\lambda}}>0$. \end{theorem} The shifted action of the affine Weyl group will be crucial in studying quantum dimensions as the following results show. \begin{theorem}\label{6:Sfixed0} For $\hat{\lambda}\in \hat{P}^{k}$ and $w\in W$, $\mathcal{D}_{w\cdot \hat{\lambda}}=(-1)^{\ell(w)}\mathcal{D}_{\hat{\lambda}}$. If $w\in W$ is an element of odd signature and $w\cdot \hat{\lambda}=\hat{\lambda}$, then $\mathcal{D}_{\hat{\lambda}}=0$. \end{theorem} \begin{theorem}\label{alcoverep} If $\mathcal{D}_{\hat{\lambda}}\neq 0$ for $\hat{\lambda}\in \hat{P}^{k}$, then we can find a unique element $ \hat{\lambda}'\in \hat{P}_{+}^{k}$ such that $\hat{\lambda}'=w\cdot \hat{\lambda}$ for some $w\in W$. \end{theorem} We now look at the role of the symmetry of the extended Dynkin diagram. \begin{theorem} \label{affinesym} If $\hat{\lambda}_1$ and $\hat{\lambda}_2\in \hat{P}^{k}$ are conjugate by an automorphism of the extended Dynkin diagram, then $\mathcal{D}_{\hat{\lambda}_1}=\mathcal{D}_{\hat{\lambda}_2}$. \end{theorem} \begin{corollary} \label{kpluszero} Let $\omega_i$ be a fundamental weight such that $\hat{\omega}_{i}$ is conjugate to $\hat{\omega}_{0}$ by an automorphism of the extended Dynkin diagram. If $\hat{\lambda}=k\hat{\omega}_i \in \hat{P}^{k}$, then $\mathcal{D}_{\hat{\lambda}}=1$. \end{corollary} \section{Preliminary computations}\label{sec:comp} From now on we will assume that $X=D_r$ and $$z^{(a)}_{m}=Q^{(a)}_{m}(\frac{\rho}{h+k})= \sum_{\omega\in \Omega^{(a)}_{m}} \mathcal{D}_{\hat{\omega}}$$ where $\Omega^{(a)}_{m}$ denotes the set of elements appearing in the sum (\ref{6:qdecompD}). \begin{proposition}\label{tipsprop} Let $a\in \{1,r-1,r\}$. The following properties hold : \begin{align} &z^{(a)}_{m}>0 \quad \text{for } 0\le m \le k, \label{1pos}\\ &z^{(a)}_m =z^{(a)}_{k-m} \quad \text{for }\, 0 < m < k ,\label{1sym}\\ &z^{(a)}_{k}=1. \label{1unit} \end{align} \end{proposition} \begin{proof} Note that $z^{(a)}_{m}=\mathcal{D}_{(k-m)\hat{\omega}_0+m\hat{\omega}_a}$. The inequality (\ref{1pos}) follows from Theorem \ref{6:qdimpos}. By Theorem \ref{affinesym}, we have $\mathcal{D}_{(k-m)\hat{\omega}_0+m\hat{\omega}_a}=\mathcal{D}_{m\hat{\omega}_0+(k-m)\hat{\omega}_a}$ and it implies (\ref{1sym}). (\ref{1unit}) is a consequence of Corollary \ref{kpluszero}. \end{proof} \begin{proposition}\label{1prop} The following properties hold : \begin{align} &z^{(1)}_{k+j}=0 \qquad \text{for }\, 1 \le j \le h -1 \label{1zer0},\\ &z^{(1)}_{k+h}=1. \label{1reunit} \end{align} \end{proposition} \begin{proof} To prove (\ref{1zer0}), we use the product formula (\ref{6:qdimdef}) for the quantum dimension. First note that for each integer $l$ such that $1 \le l \le h-1=2r-1$, there exists a positive root $\alpha$ such that $\operatorname{ht} \alpha=l$ and $\alpha-\alpha_1 \ge 0$. Moreover, the number of such roots is exactly $h=2r-2$. In the product (\ref{6:qdimdef}) for $\mathcal{D}_{(k-m)\hat{\omega}_0+m\hat{\omega}_1}$, only those roots may contribute in a non-trivial way as $$z^{(1)}_{m}=\mathcal{D}_{(k-m)\hat{\omega}_0+m\hat{\omega}_1}=\prod_{\substack{\alpha>0 \\ \alpha-\alpha_1\ge 0}}\frac{\sin\frac{\pi (\operatorname{ht}\alpha+m)}{h+k}}{\sin\frac{\pi \operatorname{ht}\alpha}{h+k}}.$$ Since $\{\operatorname{ht} \alpha+m| \alpha>0, \alpha-\alpha_1\ge 0\}$ is the same set as $\{n\in \mathbb{Z}| 1+m \leq n \leq (h-1)+m\}$, one can find a positive root $\alpha$ such that $\operatorname{ht} \alpha+m=h+k$ when $k+1 \le m \le k+(h-1)$. This proves (\ref{1zer0}). We now turn to (\ref{1reunit}). If $m=h+k$, then $$ \mathcal{D}_{(k-m)\hat{\omega}_0+m\hat{\omega}_1}=\prod_{\substack{\alpha>0 \\ \alpha-\alpha_1\ge 0}}\frac{\sin\frac{\pi(\operatorname{ht}\alpha+h+k)}{h+k}}{\sin\frac{\pi \operatorname{ht}\alpha}{h+k}}=\prod_{\substack{\alpha>0 \\ \alpha-\alpha_1\ge 0}}\frac{-\sin\frac{\pi \operatorname{ht}\alpha}{h+k}}{\sin\frac{\pi \operatorname{ht}\alpha}{h+k}}. $$ Since this product is over $h=2r-2$ terms, the final product equals 1 and it proves $z^{(1)}_{k+h}=1$. \end{proof} In the rest of the section, we will prove that for $2\le a \le r-2$, $z^{(a)}_{s}=z^{(a)}_{s+1}$ when $k$ is odd and $z^{(a)}_{s-1}=z^{(a)}_{s+1}$ when $k$ is even where $s=\lfloor\frac{k}{2}\rfloor$. If $2\le a \le r-2$, we will denote the element $k_a\hat{\omega}_{a}+k_{a-2}\hat{\omega}_{a-2}+\cdots + k_{2} \hat{\omega}_{2}+k_{0} \hat{\omega}_{0}\in \hat{P}^k$ by $(k_a,k_{a-2},\cdots, k_2, k_{0})$ when $a$ is even and $k_{a}\hat{\omega}_{a}+k_{a-2}\hat{\omega}_{a-2}+\cdots + k_{1} \hat{\omega}_{1} + k_{0} \hat{\omega}_{0}\in \hat{P}^k$ by $(k_a,k_{a-2},\cdots, k_1, k_{0})$ when $a$ is odd. Then we can write $$z^{(a)}_{m}=\sum_{\hat{\omega}\in \hat{\Omega}^{(a)}_{m}} \mathcal{D}_{\hat{\omega}}$$ where \begin{equation} \hat{\Omega}^{(a)}_{m}= \left\{(k_a,k_{a-2},\cdots, k_2, k_{0})\in \hat{P}^k\mid \begin{array}{ll} k_a+k_{a-2}+\cdots+k_2 \le m\\ k_a,k_{a-2},\cdots, k_2 \in \mathbb{Z}_{\geq 0} \\ \end{array} \right\} \end{equation} for even $a$ such that $2\le a \le r-2$ and \begin{equation} \hat{\Omega}^{(a)}_{m}= \left\{(k_a,k_{a-2},\cdots, k_1, k_{0})\in \hat{P}^k\mid \begin{array}{ll} k_a+k_{a-2}+\cdots+k_1=m\\ k_a,k_{a-2},\cdots,k_1 \in \mathbb{Z}_{\geq 0} \\ \end{array} \right\} \end{equation} for odd $a$ such that $2\le a \le r-2$. Assume that the level $k$ is odd and $s=\frac{k-1}{2}$. \begin{prop}\label{evence} If $a$ is even and $2\le a \le r-2$, then $z^{(a)}_{s}=z^{(a)}_{s+1}$. \end{prop} \begin{proof} Note that $\hat{\Omega}^{(a)}_{s}\subseteq \hat{\Omega}^{(a)}_{s+1}$ and $$ \hat{\Omega}^{(a)}_{s+1} \setminus \hat{\Omega}^{(a)}_{s}= \left\{(k_a,k_{a-2},\cdots, k_2, k_{0})\in \hat{P}^k\mid \begin{array}{ll} k_a+k_{a-2}+\cdots+k_2 = s+1 \\ k_a,k_{a-2},\cdots, k_2 \in \mathbb{Z}_{\geq 0} \\ \end{array} \right\}. $$ If $\hat{\omega}=(k_a,k_{a-2},\cdots, k_2, k_{0})\in \hat{\Omega}^{(a)}_{s+1} \setminus \hat{\Omega}^{(a)}_{s}$, then $k_0=-1$ because $$k_0+2(k_a+k_{a-2}\cdots+k_2)=k=2s+1.$$ So for any $\hat{\omega}\in \hat{\Omega}^{(a)}_{s+1} \setminus \hat{\Omega}^{(a)}_{s}$, $\mathcal{D}_{\hat\omega}=0$ since $s_0\cdot \hat\omega=\hat\omega$. Thus $z^{(a)}_{s+1}=z^{(a)}_{s}$. \end{proof} \begin{prop}\label{oddce} If $a$ is odd and $2\le a \le r-2$, then $z^{(a)}_{s}=z^{(a)}_{s+1}$. \end{prop} \begin{proof} Let $\hat{\omega}=(k_a,k_{a-2},\cdots, k_1, k_{0})\in \hat{\Omega}^{(a)}_{s+1}$. The condition \begin{equation}\label{2s1} k_0+k_1+2(k_a+k_{a-2}+\cdots+k_3)=k_0+k_1+2(s+1-k_1)=k=2s+1 \end{equation} implies $k_0=k_1-1$. If $k_1=0$, then $k_0=-1$ and $\mathcal{D}_{\hat\omega}=0$ since $s_0\cdot \hat\omega=\hat\omega$. Let $$ (\hat{\Omega}^{(a)}_{s+1})'= \left\{(k_a,k_{a-2},\cdots, k_1, k_{0})\in \hat{P}^k\mid \begin{array}{ll} k_a+k_{a-2}+\cdots+k_1=s+1 \\ k_a,k_{a-2},\cdots,k_3 \in \mathbb{Z}_{\geq 0},k_1\ge 1 \\ \end{array} \right\} $$ and $$ \hat{\Omega}^{(a)}_{s}= \left\{(k_a,k_{a-2},\cdots, k_1, k_{0})\in \hat{P}^k\mid \begin{array}{ll} k_a+k_{a-2}+\cdots+k_1=s\\ k_a,k_{a-2},\cdots,k_1 \in \mathbb{Z}_{\geq 0} \\ \end{array} \right\}. $$ Let us construct a bijection between $(\hat{\Omega}^{(a)}_{s+1})'$ and $\hat{\Omega}^{(a)}_{s}$. Define a map from $(\hat{\Omega}^{(a)}_{s+1})'$ to $\hat{\Omega}^{(a)}_{s}$ by \begin{equation}\label{map01} (k_a,k_{a-2},\cdots, k_1, k_{0}) \mapsto (k_a,k_{a-2},\cdots, k_0, k_1). \end{equation} To see that the map is well-defined, note that if $(k_a,k_{a-2},\cdots, k_1, k_{0})\in (\hat{\Omega}^{(a)}_{s+1})'\subset \hat{\Omega}^{(a)}_{s+1}$, then (\ref{2s1}) implies $k_0=k_1-1\ge 0$. Since $$ k_a+k_{a-2}+\cdots+k_3+ k_{0}=k_a+k_{a-2}+\cdots+k_3+ (k_{1}-1)=s, $$ we have $(k_a,k_{a-2},\cdots, k_0, k_1)\in \hat{\Omega}^{(a)}_{s}$. The map (\ref{map01}) is injective since $k_0=k_1-1$. Conversely, any element $(k_a,k_{a-2},\cdots, k_1, k_{0})\in \hat{\Omega}^{(a)}_{s}$ satisfies $$k_0+k_1+2(s-k_1)=k_0-k_1+2s=k=2s+1.$$ Thus $k_0=k_1+1\ge 1$ which shows that (\ref{map01}) is surjective. We thus have proved that (\ref{map01}) is a bijection between $(\hat{\Omega}^{(a)}_{s+1})'$ and $\hat{\Omega}^{(a)}_{s}$. By Theorem \ref{affinesym}, we have $\mathcal{D}_{(k_a,k_{a-2},\cdots, k_1, k_{0})}=\mathcal{D}_{(k_a,k_{a-2},\cdots, k_0, k_1)}$. This proves our assertion. \end{proof} We now assume that the level $k$ is even and $s=\frac{k}{2}$. \begin{lemma}\label{koddm4} Let $a$ be even and $2\le a \le r-2$. If $\hat\omega=(k_a,k_{a-2},\cdots, k_2, -2)\in \hat{P}^k$ satisfies $k_a+k_{a-2}+\cdots+k_2=s+1$ and $k_2=0$, then $\mathcal{D}_{\hat\omega}=0$. \end{lemma} \begin{proof} It is easy to check that $(s_0s_2s_0)\cdot \hat\omega=\hat\omega$. Theorem \ref{6:Sfixed0} now gives the desired conclusion. \end{proof} \begin{prop}\label{evence2} If $a$ is even and $2\le a \le r-2$, then $z^{(a)}_{s-1}=z^{(a)}_{s+1}$. \end{prop} \begin{proof} Recall that $$ \hat{\Omega}^{(a)}_{s-1}= \left\{(k_a,k_{a-2},\cdots, k_2, k_{0})\in \hat{P}^k\mid \begin{array}{ll} k_a+k_{a-2}+\cdots+k_2 \le s-1\\ k_a,k_{a-2},\cdots, k_2 \in \mathbb{Z}_{\geq 0} \\ \end{array} \right\} $$ and $$ \hat{\Omega}^{(a)}_{s+1}= \left\{(k_a,k_{a-2},\cdots, k_2, k_{0})\in \hat{P}^k\mid \begin{array}{ll} k_a+k_{a-2}+\cdots+k_2 \le s+1\\ k_a,k_{a-2},\cdots, k_2 \in \mathbb{Z}_{\geq 0} \\ \end{array} \right\}. $$ Let us define three disjoint subsets $R,S$ and $T$ of $\hat{\Omega}^{(a)}_{s+1}$ by \begin{align} &R=\{(k_a,k_{a-2},\cdots, k_2, k_0)\in \hat{\Omega}^{(a)}_{s+1}|k_a+k_{a-2}+\cdots+k_2=s+1, k_2 = 0\}, \notag\\ &S=\{(k_a,k_{a-2},\cdots, k_2, k_0)\in \hat{\Omega}^{(a)}_{s+1}|k_a+k_{a-2}+\cdots+k_2=s+1, k_2 \ge 1\}, \notag\\ &T=\{(k_a,k_{a-2},\cdots, k_2, k_0)\in \hat{\Omega}^{(a)}_{s+1}|k_a+k_{a-2}+\cdots+k_2=s\}. \notag \end{align} For $\hat{\omega}\in R$, $\mathcal{D}_{\hat\omega}=0$ by Lemma \ref{koddm4} and so $\sum_{\hat{\omega}\in R} \mathcal{D}_{\hat{\omega}}=0$. We now want to prove $\sum_{\hat{\omega}\in S\cup T} \mathcal{D}_{\hat{\omega}}=0$. For $(k_a,k_{a-2},\cdots, k_2, k_{0})\in \hat{P}^k$, we have $k_0+2(k_a+k_{a-2}\cdots+k_2)=k=2s$. If $(k_a,k_{a-2},\cdots, k_2, k_0)\in S$, then $k_0=-2$. For $(k_a,k_{a-2},\cdots, k_2,k_0)\in T$, we have $k_0=0$. We have a bijection between $S$ and $T$ since $$s_0\cdot (k_a,k_{a-2},\cdots, k_2, -2)=(k_a,k_{a-2},\cdots, k_2-1, 0).$$ By Theorem \ref{6:Sfixed0}, $\sum_{\hat{\omega}\in S\cup T} \mathcal{D}_{\hat{\omega}}=0$. Consequently, $$z^{(a)}_{s+1}=\sum_{\hat{\omega}\in \hat{\Omega}^{(a)}_{s+1}} \mathcal{D}_{\hat{\omega}}=\sum_{\hat{\omega}\in \hat{\Omega}^{(a)}_{s+1}\setminus (R\cup S\cup T)} \mathcal{D}_{\hat{\omega}}.$$ From $\hat{\Omega}^{(a)}_{s-1}\subset \hat{\Omega}^{(a)}_{s+1}$ and $\hat{\Omega}^{(a)}_{s-1}=\hat{\Omega}^{(a)}_{s+1}\setminus (R\cup S\cup T)$, we obtain $z^{(a)}_{s+1}=z^{(a)}_{s-1}$. \end{proof} \begin{lemma}\label{koddm5} Let $a$ be odd and $2\le a \le r-2$. If $\hat\omega=(k_a,k_{a-2},\cdots,1, -1)\in \hat{P}^k$ or $\hat\omega=(k_a,k_{a-2},\cdots, 0, -2)\in \hat{P}^k$, then $\mathcal{D}_{\hat\omega}=0$. \end{lemma} \begin{proof} For $\hat\omega=(k_a,k_{a-2},\cdots,1, -1)\in \hat{P}^k$, it is easy to see that $s_0\cdot \hat\omega=\hat\omega$. For $\hat\omega=(k_a,k_{a-2},\cdots, 0, -2)\in \hat{P}^k$, we can show $(s_0s_2s_0)\cdot \hat\omega=\hat\omega$. The lemma follows from Theorem \ref{6:Sfixed0}. \end{proof} \begin{prop}\label{oddce2} If $a$ is odd and $2\le a \le r-2$, then $z^{(a)}_{s-1}=z^{(a)}_{s+1}$ \end{prop} \begin{proof} For any $\hat{\omega}=(k_a,k_{a-2},\cdots, k_1, k_0)\in \hat{\Omega}^{(a)}_{s+1}$ with $k_1=0$ or $k_1=1$, $\mathcal{D}_{\hat\omega}=0$ by Lemma \ref{koddm5}. Let us define $(\hat{\Omega}^{(a)}_{s+1})'$ by $$\left\{(k_a,k_{a-2},\cdots, k_1, k_{0})\in \hat{P}^k\mid \begin{array}{ll} k_a+k_{a-2}+\cdots+k_1=s+1 \\ k_a,k_{a-2},\cdots,k_3 \in \mathbb{Z}_{\geq 0},k_1\ge 2 \\ \end{array} \right\}. $$ Then we can write $z^{(a)}_{s+1}=\sum_{\hat{\omega}\in (\hat{\Omega}^{(a)}_{s+1})'} \mathcal{D}_{\hat{\omega}}$. Let us define a map from $(\hat{\Omega}^{(a)}_{s+1})'$ to $\hat{\Omega}^{(a)}_{s-1}$ by \begin{equation}\label{s111} (k_a,k_{a-2},\cdots, k_1, k_{0}) \mapsto (k_a,k_{a-2},\cdots, k_0, k_1). \end{equation} For $(k_a,k_{a-2},\cdots, k_1, k_{0})\in (\hat{\Omega}^{(a)}_{s+1})'$, $$k_0+k_1+2(k_a+k_{a-2}+\cdots+k_3)=k_0+k_1+2(s+1-k_1)=k=2s.$$ Hence $k_0=k_1-2\ge 0$ and it shows that $(k_a,k_{a-2},\cdots, k_0, k_1)\in \hat{\Omega}^{(a)}_{s-1}$ and thus the map (\ref{s111}) is well-defined. It is clear that this is injective. Conversely, any element $(k_a,k_{a-2},\cdots, k_1, k_{0})\in \hat{\Omega}^{(a)}_{s-1}$ satisfies $$k_0+k_1+2(s-1-k_1)=k_0-k_1+2s-2=2s.$$ Thus $k_0=k_1+2\ge 2$ and it proves that (\ref{s111}) is surjective and thus bijective. By Theorem \ref{affinesym}, $\mathcal{D}_{(k_a,k_{a-2},\cdots, k_1, k_0)}=\mathcal{D}_{(k_a,k_{a-2},\cdots, k_0, k_1)}$ and it proves our proposition. \end{proof} \section{proof of the main theorem}\label{sec:mainpf} In this section, we prove our main theorem using the results obtained in the previous section. \begin{lemma}\label{zeq} Let $\mathbf{w}=(w^{(a)}_{m})$ be a solution of the level $k$ restricted $Q$-system such that $w^{(a)}_{m}\neq 0$ for $0\leq m\leq k$ and $a\in I$. If $w^{(a)}_{1}=Q^{(a)}_{1}$ for any $a\in I$ and $\{ Q^{(a)}_{m}|a\in I, m\in \mathbb{Z}_{\geq 0} \}$ satisfies the unrestricted $Q$-system, then $w^{(a)}_{m}=Q^{(a)}_{m}$ for $0\leq m\leq k$ and $a\in I$. In particular, $Q^{(a)}_{k}=1$. \end{lemma} \begin{proof} This is a direct consequence of the recursion (\ref{4:Qsys}) $$Q^{(a)}_{m+1}=\frac{(Q^{(a)}_{m})^2-\prod _{b\in I} (Q^{(b)}_{i})^{\mathcal{I}(X)_{ab}}}{Q^{(a)}_{m-1}}.$$ \end{proof} \begin{theorem}\label{main1} For all $a\in I$, the following properties hold : \begin{enumerate} \item (positivity) $z^{(a)}_{m}>0$ for $0\leq m \leq k$, \item (symmetry) $z^{(a)}_{m}=z_{k-m}^{(a)}$ for $1\leq m \leq k-1$, \item (unit boundary condition) $z^{(a)}_{k}=1$, \item (unimodality) $z_{m-1}^{(a)}<z^{(a)}_{m}$ for $1\le m \le \lfloor\frac{k}{2}\rfloor$. \end{enumerate} \end{theorem} \begin{proof} To prove the unit boundary condition, we divide the argument into two cases when $k$ is odd and $k$ is even. Assume first that $k$ is odd and $s=\frac{k-1}{2}$. For $0\leq m\leq k$ and $a\in I$, let us define $w^{(a)}_{m}$ by $$ w^{(a)}_{m}=\left\{ \begin{array}{ll} z^{(a)}_{m} & 0\leq m\leq s\\ z^{(a)}_{k-m}& s<m\leq k\\ \end{array} \right.. $$ Since we have $z^{(a)}_{s}=z^{(a)}_{s+1}$ for any $a\in I$ by Proposition \ref{tipsprop}, \ref{evence} and \ref{oddce}, $\mathbf{w}=(w^{(a)}_{m})$ must be a solution of the level $k$ restricted $Q$-system. Since $w^{(a)}_{1}=z^{(a)}_{1}$ and $w^{(a)}_{m}>0$ for all $0\leq m\leq k$, we can conclude that $z^{(a)}_{m}=w^{(a)}_{m}$ for $0\leq m\leq k$ by Lemma \ref{zeq}. Especially, $z^{(a)}_{k}=1$ for any $a\in I$. Assume that $k$ is even and let $s=\frac{k}{2}$. For $0\leq m\leq k$ and $a\in I$, let \begin{equation} w^{(a)}_{m}=\left\{ \begin{array}{ll} z^{(a)}_{m} & 0\leq m\leq s\\ z^{(a)}_{k-m}& s<m\leq k\\ \end{array} \right.. \end{equation} Since we have $z^{(a)}_{s-1}=z^{(a)}_{s+1}$ for $a\in I$ by Proposition \ref{tipsprop}, \ref{evence2} and \ref{oddce2}, $\mathbf{w}=(w^{(a)}_{m})$ is a solution of the level $k$ restricted $Q$-system. By the same argument as above, we can conclude that the unit boundary condition $z^{(a)}_{k}=1$ holds for any $a\in I$. The properties of positivity and symmetry can be easily obtained from the definition for $w^{(a)}_{m}$. Now we know that $\mathbf{w}=\mathbf{z}=(z^{(a)}_{m})$ is the positive solution of the level $k$ restricted $Q$-system characterized in Theorem \ref{5:Quni} and it follows that $z_{m-1}^{(a)}<z^{(a)}_{m}$ for $1\le m \le \lfloor\frac{k}{2}\rfloor$. \end{proof} Now we prove $z^{(a)}_{k+j}=0$ for any $a\in I$ and $1 \le j \le h -1$. \begin{proposition}\label{kpl10} For all $a\in I$, $z^{(a)}_{k+1}=0$. \end{proposition} \begin{proof} Since $z^{(a)}_{k}=1$ and $z^{(a)}_{k-1}\neq 0$ for all $a\in I$ by Theorem \ref{main1}, the recursion (\ref{4:Qsys}) $$(z^{(a)}_{k})^2=\prod _{b\in I} (z^{(b)}_{k})^{\mathcal{I}(X)_{ab}}+z^{(a)}_{k-1}z^{(a)}_{k+1}$$ implies $z^{(a)}_{k+1}=0$. \end{proof} \begin{lemma}\label{arr0} Let $\{ Q^{(a)}_{m}|a\in I, m\in \mathbb{Z}_{\geq 0} \}$ be a solution of the unrestricted $Q$-system. The following condition $$\left[\begin{array}{ccc} Q^{(a-1)}_{m-1} & Q^{(a)}_{m-1} & Q^{(a+1)}_{m-1} \\ Q^{(a-1)}_{m} & Q^{(a)}_{m} & Q^{(a+1)}_{m} \\ Q^{(a-1)}_{m+1} & Q^{(a)}_{m+1} & Q^{(a+1)}_{m+1} \end{array}\right]=\left[\begin{array}{ccc} * & 0 & * \\ 0 & Q^{(a)}_{m} & * \\ * & * & * \end{array}\right]$$ implies $Q^{(a)}_{m}=0$ for $2\leq a\leq r-2$ where $*$ denotes an arbitrary number. Similarly, the condition $$\left[\begin{array}{ccc} Q^{(r-2)}_{i-1} & Q^{(r-1)}_{i-1} & Q^{(r)}_{i-1} \\ Q^{(r-2)}_{i} & Q^{(r-1)}_{i} & Q^{(r)}_{i} \\ Q^{(r-2)}_{i+1} & Q^{(r-1)}_{i+1} & Q^{(r)}_{i+1} \end{array}\right]=\left[\begin{array}{ccc} * & 0 & 0 \\ 0 & Q^{(r-1)}_{m} & Q^{(r)}_{m} \\ * & * & * \end{array}\right]$$ implies $Q^{(r-1)}_{m}=Q^{(r)}_{m}=0$. \end{lemma} \begin{proof} This is again a consequence of the recursion (\ref{4:Qsys}) for $Q^{(a)}_{m}$: $$(Q^{(a)}_{m})^2=\prod _{b\in I} (Q^{(b)}_{i})^{\mathcal{I}(X)_{ab}}+Q^{(a)}_{m-1}Q^{(a)}_{m+1}.$$ In both cases, we obtain $(Q^{(a)}_{m})^2=0$. \end{proof} \begin{theorem}\label{main2} $z^{(a)}_{k+j}=0$ for any $a\in I$ and $1 \le j \le h -1$. \end{theorem} \begin{proof} Recall (\ref{1zer0}) that $z^{(1)}_{k+j}=0$ for $1 \le j \le h -1$. Since $z^{(a)}_{k+1}=0$ for any $a\in I$ by Proposition \ref{kpl10}, we get $z^{(a)}_{k+2}=0$ by applying Lemma \ref{arr0}. Repeated application of Lemma \ref{arr0} enables us to prove $z^{(a)}_{k+j}=0$ for any $a\in I$ and $1 \le j \le h -1$. \end{proof} For the following lemma, let us set $Q^{(0)}_{m}=1$ for convenience. \begin{lemma}\label{arr1} Let $\{ Q^{(a)}_{m}|a\in I, m\in \mathbb{Z}_{\geq 0} \}$ be a solution of the unrestricted $Q$-system. The following condition $$ \left[\begin{array}{cccc} Q^{(a-2)}_{m-1} & Q^{(a-1)}_{m-1} & Q^{(a)}_{m-1} \\ Q^{(a-2)}_{m} & Q^{(a-1)}_{m} & Q^{(a)}_{m} \\ Q^{(a-2)}_{m+1} & Q^{(a-1)}_{m+1} & Q^{(a)}_{m+1} \end{array}\right]= \left[\begin{array}{cccc} * & 0 & 0 \\ 1 & 1 & Q^{(a)}_{m} \\ * & * & * \end{array}\right]$$ implies $Q^{(a)}_{m}=1$ for $2\leq a\leq r-2$ where $*$ denotes an arbitrary number. \end{lemma} \begin{proof} Let us look at (\ref{4:Qsys}) for $Q^{(a-1)}_{m}$, $$(Q^{(a-1)}_{m})^2=Q^{(a-2)}_{m}Q^{(a)}_{m}+Q^{(a-1)}_{m-1}Q^{(a-1)}_{m+1}.$$ Under the condition stated above, we obtain $1^2=1\cdot Q^{(a)}_{m}+0$. This proves $Q^{(a)}_{m}=1$. \end{proof} \begin{theorem}\label{main3} $z^{(a)}_{k+h}=1$ for $1\le a\le r-2$ and $z^{(r-1)}_{k+h}=z^{(r)}_{k+h}=\pm 1$. \end{theorem} \begin{proof} Since $z^{(1)}_{k+h}=1$ by (\ref{1reunit}) and $z^{(a)}_{k+h-1}=0$ for any $a\in I$ by Theorem \ref{main2}, we get $z^{(a)}_{k+h}=1$ for $1\le a \le r-2$ by applying Lemma \ref{arr1}. The recursions (\ref{4:Qsys}) for $z^{(r-2)}_{k+h}$, $z^{(r-1)}_{k+h}$ and $z^{(r)}_{k+h}$ give the following system of equations $$ \left\{ \begin{array}{lll} &(z^{(r-2)}_{k+h})^2=z^{(r-1)}_{k+h}z^{(r)}_{k+h}\\ &(z^{(r-1)}_{k+h})^2=z^{(r-2)}_{k+h}\\ &(z^{(r)}_{k+h})^2=z^{(r-2)}_{k+h} \end{array} \right.. $$ Then $z^{(r-2)}_{k+h}=1$ implies $z^{(r-1)}_{k+h}=z^{(r)}_{k+h}=\pm 1$. \end{proof} \begin{remark} Analysis similar to that in the proof of Proposition \ref{1prop} using the product formula (\ref{6:qdimdef}) for the quantum dimension can be used to show $z^{(r-1)}_{k+h}=z^{(r)}_{k+h}=1$ when $r\equiv 0,1 \pmod 4$ and $z^{(r-1)}_{k+h}=z^{(r)}_{k+h}=-1$ when $r\equiv 2,3 \pmod 4$. \end{remark} \begin{example} Let $X=D_{5}$ and $k=4$. We express $z_m^{(a)}$ in terms of quantum dimensions for $a\in I$ and $0\le m \le h+k=12$. For $a=1,4,5$, we get $$ \begin{bmatrix} z_0^{(1)} & z_0^{(4)} & z_0^{(5)} \\ z_1^{(1)} & z_1^{(4)} & z_1^{(5)} \\ z_2^{(1)} & z_2^{(4)} & z_2^{(5)} \\ z_3^{(1)} & z_3^{(4)} & z_3^{(5)} \\ z_4^{(1)} & z_4^{(4)} & z_4^{(5)} \\ z_5^{(1)} & z_5^{(4)} & z_5^{(5)} \\ \vdots & \vdots & \vdots \\ z_{11}^{(1)} & z_{11}^{(4)} & z_{11}^{(5)} \\ z_{12}^{(1)} & z_{12}^{(4)} & z_{12}^{(5)} \end{bmatrix} = \begin{bmatrix} \mathcal{D}_{4 \hat{\omega }_0} & \mathcal{D}_{4 \hat{\omega }_0} & \mathcal{D}_{4 \hat{\omega }_0} \\ \mathcal{D}_{3 \hat{\omega }_0+\hat{\omega }_1} & \mathcal{D}_{3 \hat{\omega }_0+\hat{\omega }_4} & \mathcal{D}_{3 \hat{\omega }_0+\hat{\omega }_5} \\ \mathcal{D}_{2 \hat{\omega }_0+2 \hat{\omega }_1} & \mathcal{D}_{2 \hat{\omega }_0+2 \hat{\omega }_4} & \mathcal{D}_{2 \hat{\omega }_0+2 \hat{\omega }_5} \\ \mathcal{D}_{\hat{\omega }_0+3 \hat{\omega }_1} & \mathcal{D}_{\hat{\omega }_0+3 \hat{\omega }_4} & \mathcal{D}_{\hat{\omega }_0+3 \hat{\omega }_5} \\ \mathcal{D}_{4 \hat{\omega }_1} & \mathcal{D}_{4 \hat{\omega }_4} & \mathcal{D}_{4 \hat{\omega }_5} \\ 0 & 0 & 0 \\ \vdots & \vdots & \vdots \\ 0 & 0 & 0 \\ \mathcal{D}_{4 \hat{\omega }_1} & \mathcal{D}_{4 \hat{\omega }_4} & \mathcal{D}_{4 \hat{\omega }_5} \end{bmatrix}. $$ For $a=2,3$, we have $$ \begin{bmatrix} z_0^{(2)} & z_0^{(3)} \\ z_1^{(2)} & z_1^{(3)} \\ z_2^{(2)} & z_2^{(3)} \\ z_3^{(2)} & z_3^{(3)} \\ z_4^{(2)} & z_4^{(3)} \\ z_5^{(2)} & z_5^{(3)} \\ \vdots & \vdots \\ z_{11}^{(2)} & z_{11}^{(3)} \\ z_{12}^{(2)} & z_{12}^{(3)} \end{bmatrix}= \begin{bmatrix} \mathcal{D}_{4 \hat{\omega }_0} & \mathcal{D}_{4 \hat{\omega }_0} \\ \mathcal{D}_{4 \hat{\omega }_0}+\mathcal{D}_{2 \hat{\omega }_0+\hat{\omega }_2} & \mathcal{D}_{3 \hat{\omega }_0+\hat{\omega }_1}+\mathcal{D}_{2 \hat{\omega }_0+\hat{\omega }_3} \\ \mathcal{D}_{4 \hat{\omega }_0}+\mathcal{D}_{2 \hat{\omega }_2}+\mathcal{D}_{2 \hat{\omega }_0+\hat{\omega }_2} & \mathcal{D}_{2 \hat{\omega }_0+2 \hat{\omega }_1}+\mathcal{D}_{2 \hat{\omega }_3}+\mathcal{D}_{\hat{\omega }_0+\hat{\omega }_1+\hat{\omega }_3} \\ \mathcal{D}_{4 \hat{\omega }_0}+\mathcal{D}_{2 \hat{\omega }_0+\hat{\omega }_2} & \mathcal{D}_{\hat{\omega }_0+3 \hat{\omega }_1}+\mathcal{D}_{2 \hat{\omega }_1+\hat{\omega }_3} \\ \mathcal{D}_{4 \hat{\omega }_0} & \mathcal{D}_{4 \hat{\omega }_1} \\ 0 & 0 \\ \vdots & \vdots\\ 0 & 0 \\ \mathcal{D}_{4 \hat{\omega }_0} & \mathcal{D}_{4 \hat{\omega }_1} \end{bmatrix}.$$ This expression is obtained by applying the shifted affine Weyl group action together with Theorem \ref{6:Sfixed0} and \ref{alcoverep}. \end{example} \section*{Acknowledgements} This work is an outgrowth of the author's doctoral research and he would like to thank his advisor Richard E. Borcherds. He is also grateful to Tomoki Nakanishi for helpful discussions at the MSRI workshop on Cluster Algebras and Commutative Algebra. The work is partially supported by Samsung Scholarship.
\section{Introduction} \label{sc:intro} The detection of hot atomic hydrogen in the upper atmosphere of HD209458b \citep{vidalmadjar03,vidalmadjar04} has inspired numerous attempts to model physical and chemical processes in highly irradiated atmospheres, including rapid escape as one of the most challenging aspects. Subsequent detection of heavy atoms and ions \citep{vidalmadjar04,linsky10} point out the need for more complex models that include the chemistry associated with these species as well as the collision coupling between them and the major species. Indeed, close-in extrasolar planets offer a natural laboratory to constrain the theory of rapid escape, including hydrodynamic escape. This is important because aspects of the theory are controversial, and yet rapid escape is believed to have played a role in shaping the early evolution of the atmospheres in the solar system \citep[e.g.,][]{zahnle86,hunten87}. Escape may also be a crucial factor in determining atmospheric conditions and habitability of super-Earths and Earth-like planets around M dwarfs \citep[e.g.,][]{tarter07} that may be amenable to spectroscopic studies in the near future \citep[e.g.,][]{charbonneau09}. The basic ideas about the nature of the upper atmospheres around close-in EGPs were laid out almost as soon as the first planet, 51 Peg b \citep{mayor95}, was detected. For instance, \citet{coustenis98} argued that heating by the stellar EUV radiation and interaction with the stellar wind leads to high temperatures of several thousand Kelvins in the upper atmosphere and exosphere of close-in EGPs. They also suggested that the upper atmosphere is primarily composed of atoms and ions, and that hydrodynamic escape might be able to drag species heavier than H and He into the exosphere. At the same time, \citet{schneider98} argued that material escaping from the atmospheres of close-in EGPs would form a potentially observable comet-like tail. When \citet{vidalmadjar03,vidalmadjar04} detected the transits of HD209458b in the stellar FUV emission lines, they also argued that the planet is followed by comet-like tail of escaping hydrogen, and that hydrodynamic escape is required to drag oxygen and carbon atoms to the thermosphere. The model of \citet{yelle04,yelle06} was the first attempt to model the aeronomy and escape processes in detail and most of the assumptions in that work have been adopted by subsequent investigators. It solved the vertical equations of continuity, momentum, and energy for an escaping atmosphere, including photochemistry in the ionosphere and transfer of stellar XUV radiation. Based on a composition of hydrogen and helium, the results demonstrated that H$_2$ dissociates in the thermosphere, which at high altitudes is dominated by H and H$^+$. The model also showed that stellar heating leads to temperatures of $\sim$10,000 K in the upper atmosphere, and predicted an energy-limited mass loss rate of 4.7~$\times$~10$^7$ kg~s$^{-1}$ \citep{yelle06}. \citet{yelle04} argued that conditions beyond $\sim$3 $R_p$ were too complex and uncertain to be modeled reliably and therefore chose an upper boundary at 3 $R_p$, rather than at infinity, as adopted in early solar wind models. This led to a requirement for boundary conditions for the fluid equations at a finite radius. \citet{yelle04} required consistency between fluid and kinetic simulations, based on the well established fact that kinetic and fluid approaches provide consistent results for the escape flux \citep[e.g.,][]{lemaire73}. This led to subsonic velocities of a few km~s$^{-1}$ in his model -- although the presence of a sonic point at a higher altitude was not ruled out. Many other models for the upper atmospheres of close-in EGPs have been published \citep[e.g.,][]{lammer03,lecavelier04,jaritz05,tian05,erkaev07,garciamunoz07,schneiter07,penz08,holstrom08,murrayclay09,stone09,guo11,trammell11}. These include one-dimensional, two-dimensional, and three-dimensional models that make different assumptions regarding heating efficiency, the effect of stellar tides, photochemistry, and the escape mechanism. Despite significant differences in the temperature and velocity profiles, almost all of the existing models agree that close-in EGPs such as HD209458b are surrounded by an extended, hot thermosphere that is undergoing some form of escape. Most of the models to date concentrate on the distribution of H and H$^+$ in the upper atmosphere. \citet{garciamunoz07} developed the only model to address the presence of O and C$^+$ in the thermosphere \citep{vidalmadjar04,linsky10}. This model is otherwise similar to \citet{yelle04}, but it includes the photochemistry of heavy ions, atoms, molecules, and molecular ions. It also extends to higher altitudes, and includes the effect of substellar tidal forces and stellar wind, albeit in an approximate manner. \citet{koskinen07a,koskinen07} developed a three-dimensional model for the thermospheres of EGPs at wide orbits. They pointed out that the atmospheres of close-in EGPs do not escape hydrodynamically unless they receive enough stellar XUV energy to dissociate molecules in the EUV heating layer below the exobase. Although their results are limited to the specific case of H$_2$, they can be generalized as follows. The most important molecules H$_2$ (through the formation of H$_3^+$), CO, H$_2$O, and CH$_4$ act as strong infrared coolants in the thermosphere. High temperatures and rapid escape are only possible once these molecules are dissociated. \citet{koskinen07} showed that H$_2$ dissociates in the thermosphere of a Jupiter-type planet orbiting a Sun-like star within 0.2 AU. Once H$_2$ dissociates, it is reasonable to assume that other molecules dissociate too. At this point the pressure scale height is enhanced by a factor of $\sim$10 when H becomes the dominant species in the thermosphere and temperatures reach 10,000 K. It should be noted that a composition of H and H$^+$ with high temperatures does not guarantee that the atmosphere escapes hydrodynamically. For instance, \citet{koskinen09} showed that hydrodynamic escape is extremely unlikely to occur on a planet such as HD17156b because of its high mass and eccentric orbit. These types of results have implications on statistical studies that characterize the escape of planetary atmospheres by relying on the so-called energy-limited escape \citep[e.g.,][]{watson81,lecavelier07,sanzforcada10}. These studies often include an efficiency factor in the mass loss rate that is based on the heating efficiency of the upper atmosphere \citep[e.g.,][]{lammer09}. Unless the atmosphere is escaping rapidly, the heating efficiency could be considerably larger than the fraction of energy that actually powers escape through adiabatic cooling. Under diffusion-limited escape or in the Jeans regime the energy-limited escape rate is just an upper limit and the true escape rate can be lower. Ideally, the uncertainties in the models can be limited by detailed observations of the escaping species. At present, multiple observations are only available for HD209458b, and they reveal the presence of H, O, C$^+$, and Si$^{2+}$ at high altitudes in the thermosphere \citep{vidalmadjar03,vidalmadjar04,linsky10}. Visible and infrared observations have also revealed the presence of Na, H$_2$O, CH$_4$, and CO$_2$ in the lower atmosphere \citep{charbonneau02,knutson08,swain09}. Taken together, these observations are beginning to reveal the composition and thermal structure in the atmosphere of HD209458b. The purpose of the current paper is to characterize the density profiles of all of the detected species in the thermosphere, and to explain the presence of the heavy atoms and ions at high altitudes in the upper atmosphere. The results can be used to infer some basic properties of the atmosphere. To this end, we introduce a one-dimensional escape model for the upper atmosphere of HD209458b that includes the photochemistry of heavy atoms and ions. As pointed out above, previous models agree broadly on the qualitative nature of the thermosphere but the temperature, density, and velocity profiles predicted by them differ significantly (see Section~\ref{subsc:tempvel}). Some authors have argued that the density of H in the thermosphere is not sufficient to explain the observed transit depths \citep[see][for a review]{koskinen10}, thus lending support to alternative interpretations of the observations such as the presence of energetic neutral atoms \citep{holstrom08} or a comet-like tail of hydrogen shaped by radiation pressure \citep{vidalmadjar03}. Accurate modeling of the thermosphere is required to enable better judgment between different explanations of the observations. The differencies between previous models arise from different assumptions regarding heating rates and boundary conditions. In addition to modeling the density profiles of the detected heavy species, we have improved these aspects of the calculations in our work. For instance, the lower boundary conditions are constrained by results from a detailed photochemical model of the lower atmosphere (Lavvas et al., \textit{in preparation}). With regard to the upper boundary conditions, we demonstrate that for HD209458b the extrapolated `outflow' boundary conditions \citep[e.g.,][]{tian05} are consistent with recent results from kinetic theory \citep{volkov11a,volkov11b} as long as the upper boundary is at a sufficiently high altitude -- although uncertainties regarding the interaction of the atmosphere with the stellar wind may limit the validity of both boundary conditions. We highlight the effect of heating efficiency and stellar flux on the density and temperature profiles, and constrain the likely heating rates by using photoelectron heating efficiencies based on the results of \citet{cecchi09} and our own estimates (Section~\ref{subsc:tempvel}). As a result we provide a robust qualitative description of the density profiles, and constrain the mean temperature and velocity profile in the thermosphere. A second paper by \citet{koskinen12b} (Paper II) compares our results directly with the observations. \section{Methods} \label{sc:methods} \subsection{Hydrodynamic model} \label{subsc:hydromodel} We use a one-dimensional escape model for HD209458b ($R_p =$~1.32 $R_J$, $M_p =$~0.69 $M_J$, $a =$~0.047 AU) that is similar to the models of \citet{yelle04} and \citet{garciamunoz07}. Because such models are extensively discussed in the literature, we include only a brief overview of the model here, with the emphasis on how it differs from previous work. The model solves the one-dimensional equations of motion for an escaping atmosphere composed of several neutral and ionized species: \begin{eqnarray} \frac{\partial \rho_s}{\partial t} + \frac{1}{r^2} \frac{\partial}{\partial r} (r^2 \rho_s v)&+&\frac{1}{r^2} \frac{\partial}{\partial r} (r^2 F_s) = \sum_t R_{st} \\ \frac{\partial (\rho v)}{\partial t} + \frac{1}{r^2} \frac{\partial}{\partial r} (r^2 \rho v^2)&=&-\rho g - \frac{\partial p}{\partial r} + f_{\mu} \\ \frac{\partial (\rho E)}{\partial t} + \frac{1}{r^2} \frac{\partial}{\partial r} (r^2 \rho E v)&=&\rho Q_R - p \frac{1}{r^2} \frac{\partial}{\partial r} (r^2 v) \nonumber \\ + \frac{1}{r^2} \frac{\partial}{\partial r} \left( r^2 \kappa \frac{\partial T}{\partial r} \right)&+&\Phi_{\mu} \end{eqnarray} where $\rho_s$ is the density of species $s$, $v$ is the vertical velocity, $F_s$ is the diffusive flux of species $s$, $R_{st}$ is the net chemical source term for species $s$, $f_{\mu}$ is a force term arising from viscous acceleration, $E = c_v T$ is the specific internal energy of the gas, $Q_R$ is the specific net radiative heating rate, $\kappa$ is the coefficient of heat conduction, and $\Phi_{\mu}$ is the viscous dissipation functional \citep[e.g.,][]{oneill89}. The total density and pressure are given by $\rho = \sum_s \rho_s$ and $p = \sum_s n_s k T$, respectively, where electrons contribute to the total pressure. We assumed equal temperatures for the neutral species, ions and electrons, and calculated the electron density at each altitude from the requirement of charge neutrality. The model solves separate continuity equations for each species, but treats the atmosphere otherwise as a single fluid. The differences in the velocities of the individual species are taken into account by including the diffusive flux $F_s$. We calculated the fluxes by solving simultaneous equations for multiple species based on the diffusion equation given by \citet{chapman70} (equation 18.2,6, p.344). We also included a force term due to the ambipolar electric field given by $eE = -(1/n_e) \text{d} p_e / \text{d} r$, where the subscript $e$ refers to electrons, that can be important in highly ionized flows. The collision terms account for neutral-neutral, resonant and non-resonant ion-neutral, and Coulomb collisions. This method is in principle similar to those of \citet{yelle04} and \citet{garciamunoz07}. We verify that the single temperature and diffusion approximations are valid for HD209458b based on our results in Section~\ref{subsc:ionescape}. The model includes heat conduction and terms due to viscosity in both the momentum and energy equations. Thus the equations are consistent with the level of approximation in the Navier-Stokes (NS) equations. The NS equations themselves are a simplification of the 13-moment solution to the Boltzmann equation \citep[e.g.,][]{gombosi94} that is valid when the Knudsen number $Kn = \Lambda/L << 1$, where $\Lambda$ is the mean free path and $L$ is the typical length scale for significant changes in density or temperature. Broadly speaking, the equations are valid below the exobase, and terms due to heat conduction and viscosity gain significance as $Kn \rightarrow$~1. We note that the exobase on HD209458b is typically located at a very high altitude (see Section~\ref{subsc:tempvel}), and viscosity and heat conduction are not particularly important. We included species such as H, H$^+$, He, He$^+$, C, C$^+$, O, O$^+$, N, N$^+$, Si, Si$^+$, Si$^{2+}$, and electrons in the hydrodynamic model. We also generated simulations that included Mg, Mg$^{+}$, Na, Na$^{+}$, K, K$^+$, S, and S$^+$, but the presence of these species did not affect the density profiles of H, O, C$^+$, or Si$^{2+}$ significantly. The model includes photoionization, thermal ionization, and charge exchange between atoms and ions. The reaction rate coefficients for these processes are listed in Table~\ref{table:reactions}. Multiply charged ions were included only if the ionization potential of their parent ion was sufficiently low compared to the thermal energy and radiation field in the upper atmosphere. We note that our model also includes impact ionization by thermal electrons. In general, this can be important for species with low ionization potential such as alkali metals \citep[e.g.,][]{batygin10}, although we find photoionization to be more significant in the thermosphere (see Section~\ref{subsc:denprofs}). In order to simulate photochemistry in a numerically robust fashion, we coupled the dynamical model with the ASAD chemistry integrator developed at the University of Cambridge \citep{carver97}. In most cases we used the IMPACT integration scheme that is provided by ASAD. We did not include any molecules in the present simulations, and thus placed the lower boundary of the hydrodynamic model at $p_0 =$~1 $\mu$bar (see Section~\ref{subsc:photochemistry}). Molecular chemistry is not significant in the thermosphere, where our results agree qualitatively with \citet{garciamunoz07} despite simpler chemistry (see Section~\ref{subsc:denprofs}). This is an important result because it implies that complex molecular photochemistry does not need to be included in the models for the thermosphere. However, the chemistry of molecular ions may be important on HD209458b below the 0.1 $\mu$bar level and it needs to be studied in greater detail. \begin{table}[htbp] \tiny{ \centering \caption{Reaction rate coefficients} \begin{tabular}{@{} lcc @{}} \toprule \cmidrule( r ){1-3} Reaction & Rate (cm$^3$ s$^{-1}$) & Reference \\ \midrule P1 \ \ \ \ H$\ + \ h\nu \ \rightarrow$~H$^+ \ + \ e$ & & \citep{hummer63} \\ P2 \ \ \ \ He$\ + \ h\nu \ \rightarrow$~He$^+ \ + \ e$ & & \citep{yan98} \\ P3 \ \ \ \ O$\ + \ h\nu \ \rightarrow$~H$^+ \ + \ e$ & & \citep{verner96} \\ P4 \ \ \ \ C$\ + \ h\nu \ \rightarrow$~C$^+ \ + \ e$ & & \citep{verner96} \\ P5 \ \ \ \ N$\ + \ h\nu \ \rightarrow$~N$^+ \ + \ e$ & & \citep{verner96} \\ P6 \ \ \ \ Si$\ + \ h\nu \ \rightarrow$~Si$^+ \ + \ e$ & & \citep{verner96} \\ P7 \ \ \ \ Si$^+\ + \ h\nu \ \rightarrow$~Si$^{2+} \ + \ e$ & & \citep{verner96} \\ R1 \ \ \ \ H$^+ + e \rightarrow$~H~$ + \ h\nu$ & 4.0~$\times$~10$^{-12} (300/T_e)^{0.64}$ & \citep{storey95} \\ R2 \ \ \ \ He$^+ + e \rightarrow$~He~$ + \ h\nu$ & 4.6~$\times$~10$^{-12} (300/T_e)^{0.64}$ & \citep{storey95} \\ R3 \ \ \ \ H$\ +\ e \rightarrow$~H$^+ + e + e$ & 2.91~$\times$~10$^{-8} \left( \frac{1}{0.232+U} \right) U^{0.39} \exp(-U) \ , \ U = 13.6/E_e(eV)$ & \citep{voronov97} \\ R4 \ \ \ \ He$\ +\ e \rightarrow$~He$^+ + e + e$ & 1.75~$\times$~10$^{-8} \left( \frac{1}{0.180+U} \right) U^{0.35} \exp(-U) \ , \ U = 24.6/E_e(eV)$ & \citep{voronov97} \\ R5 \ \ \ \ H$ \ +$~He$^+ \rightarrow$~H$^+ \ + \ $He & 1.25~$\times$~10$^{-15} (300/T)^{-0.25}$ & \citep{glover07} \\ R6 \ \ \ \ H$^+ \ +$~He$\ \rightarrow$~H$ \ + \ $He$^+$ & 1.75~$\times$~10$^{-11} (300/T)^{0.75} \exp(-128,000/T)$ & \citep{glover07} \\ R7 \ \ \ \ O$\ +\ e \rightarrow$~O$^+ + e + e$ & 3.59~$\times$~10$^{-8} \left( \frac{1}{0.073+U} \right) U^{0.34} \exp(-U) \ , \ U = 13.6/E_e(eV)$ & \citep{voronov97} \\ R8 \ \ \ \ C$\ +\ e \rightarrow$~C$^+ + e + e$ & 6.85~$\times$~10$^{-8} \left( \frac{1}{0.193+U} \right) U^{0.25} \exp(-U) \ , \ U = 11.3/E_e(eV)$ & \citep{voronov97} \\ R9 \ \ \ \ O$^+ + e \rightarrow$~O~$ + \ h\nu$ & 3.25~$\times$~10$^{-12} (300/T_e)^{0.66}$ & \citep{woodall07} \\ R10 \ \ C$^+ + e \rightarrow$~C~$ + \ h\nu$ & 4.67~$\times$~10$^{-12} (300/T_e)^{0.60}$ & \citep{woodall07} \\ R11 \ \ C$^+ \ +$~H $\ \rightarrow$~C~$ + \ $ H$^+$ & 6.30~$\times$~10$^{-17} (300/T)^{-1.96} \exp(-170,000/T)$ & \citep{stancil98} \\ R12 \ \ C$ \ +$~H$^+ \ \rightarrow$~C$^+ \ + \ $ H & 1.31~$\times$~10$^{-15} (300/T)^{-0.213}$ & \citep{stancil98} \\ R13 \ \ C$ \ +$~He$^+ \ \rightarrow$~C$^+ \ + \ $ He & 2.50~$\times$~10$^{-15} (300/T)^{-1.597}$ & \citep{glover07} \\ R14 \ \ O$^+ \ +$~H $\ \rightarrow$~O~$ + \ $ H$^+$ & 5.66~$\times$~10$^{-10} (300/T)^{-0.36} \exp(8.6/T)$ & \citep{woodall07} \\ R15 \ \ O$ \ +$~H$^+ \ \rightarrow$~O$^+ \ + \ $ H & 7.31~$\times$~10$^{-10} (300/T)^{-0.23} \exp(-226.0/T)$ & \citep{woodall07} \\ R16 \ \ N$\ +\ e \rightarrow$~N$^+ + e + e$ & 4.82~$\times$~10$^{-8} \left( \frac{1}{0.0652+U} \right) U^{0.42} \exp(-U) \ , \ U = 14.5/E_e(eV)$ & \citep{voronov97} \\ R17 \ \ N$^+ + e \rightarrow$~N~$ + \ h\nu$ & 3.46~$\times$~10$^{-12} (300/T_e)^{0.608}$ & \citep{aldrovandi73} \\ R18 \ \ Si$\ +\ e \rightarrow$~Si$^+ + e + e$ & 1.88~$\times$~10$^{-7} \left( \frac{1+\sqrt{U}}{0.376+U} \right) U^{0.25} \exp(-U) \ , \ U = 8.2/E_e(eV)$ & \citep{voronov97} \\ R19 \ \ Si$^+ + e \rightarrow$~Si~$ + \ h\nu$ & 4.85~$\times$~10$^{-12} (300/T_e)^{0.60}$ & \citep{aldrovandi73} \\ R20 \ \ Si$^+ \ +\ e \rightarrow$~Si$^{2+} + e + e$ & 6.43~$\times$~10$^{-8} \left( \frac{1+\sqrt{U}}{0.632+U} \right) U^{0.25} \exp(-U) \ , \ U = 16.4/E_e(eV)$ & \citep{voronov97} \\ R21 \ \ Si$^{2+} + e \rightarrow$~Si$^+$~$ + \ h\nu$ & 1.57~$\times$~10$^{-11} (300/T_e)^{0.786}$ & \citep{aldrovandi73} \\ R22 \ \ H$^+ \ +$~Si $\ \rightarrow$~H~$ + \ $ Si$^+$ & 7.41~$\times$~10$^{-11} (300/T)^{-0.848}$ & \citep{glover07} \\ R23 \ \ He$^+ \ +$~Si $\ \rightarrow$~He~$ + \ $ Si$^+$ & 3.30~$\times$~10$^{-9}$ & \citep{woodall07} \\ R24 \ \ C$^+ \ +$~Si $\ \rightarrow$~C~$ + \ $ Si$^+$ & 2.10~$\times$~10$^{-9}$ & \citep{woodall07} \\ R25 \ \ H$ \ +$~Si$^{2+} \ \rightarrow$~H$^+ \ + \ $ Si$^+$ & 2.20~$\times$~10$^{-9} (300/T)^{-0.24}$ & \citep{kingdon96} \\ R26 \ \ H$^+ \ +$~Si$^+ \ \rightarrow$~H$ \ + \ $ Si$^{2+}$ & 7.37~$\times$~10$^{-10} (300/T)^{-0.24}$ & \citep{kingdon96} \\ \bottomrule \end{tabular} \label{table:reactions}} \end{table} The upper atmosphere is heated by stellar XUV radiation. We simulated heating and photoionization self-consistently by using the model density profiles and the UV spectrum of the average Sun. The spectrum covers wavelengths between 0.1--3000~\AA. The XUV spectrum between 0.1--1050~\AA~was generated by the SOLAR2000 model \citep{tobiska00}. It includes strong emission lines separately and weaker lines binned by 50~\AA. The Lyman~$\alpha$ line was included with a wavelength spacing of 0.5~\AA~from \citet{lemaire05} and the rest of the spectrum was taken from \citet{woods02}. We assumed that most of the Lyman $\alpha$ radiation absorbed by H is resonantly scattered and does not contribute significantly to the heating of the atmosphere. This is because the lifetime of the 2p state of H is only 1.6 ns, compared with the typical collision timescale of $\sim$1 s near the temperature peak in the thermosphere of HD209458b. References for photoabsorption cross sections of the different species are included in Table~\ref{table:reactions}. In general, we divided the incident stellar flux by a factor of 4 to account for uniform redistribution of energy around the planet. This is expected to be approximately valid in the lower thermosphere based on the three-dimensional simulations of \citet{koskinen10b}. In the extended upper thermosphere, on the other hand, radiation passes through to the night side and leaves only a small region free of direct heating \citep[e.g., see Figure 2 of][]{koskinen07}. The current model also includes heating due to photoabsorption by C, O, N, and metals. This is relatively insignificant -- although it leads to some additional heating in the lower thermosphere by FUV radiation. Heating of the thermosphere is mostly driven by photoionization and the generation of photoelectrons, although direct excitation of atoms and molecules may also play a role. Photoelectrons excite, ionize, and dissociate atoms and molecules until they lose enough energy and become thermalized i.e., share their energy with thermal electrons in Coulomb collisions. Thermal electrons share their energy with ions and eventually, the neutral atmosphere. In highly ionized atmospheres such as on HD209458b the photoelectron heating efficiency can be close to 100~\% \citep{cecchi09}, depending on the energy of the photoelectrons. We used scaled heating efficiencies that depend on photoelectron energy to estimate the net heating efficiency in the atmosphere (Section~\ref{subsc:tempvel}). Generally, we refer to two different definitions of heating efficiency in Section~\ref{subsc:tempvel} in order to highlight the effect of heating efficiency on the temperature and velocity profiles. The net heating efficiency $\eta_{\text{net}}$ is defined simply as the fraction of the absorbed stellar energy that heats the atmosphere. Photoelectron heating efficiency, on the other hand, applies to photoelectrons with energy $E_p = h \nu - I_s$ where $I_s$ is the ionization potential of species $s$ and $h \nu$ is the energy of the ionizing photon. The photoelectron heating efficiency is the fraction of $E_p$ that heats the thermosphere, and it is generally higher than $\eta_{\text{net}}$ because it does not account for recombination. The net heating efficiency is often used to calculate mass loss rates for extrasolar planets \citep[e.g.,][]{lammer09}. Therefore it is important not to confuse these two definitions of heating efficiency. We included radiative cooling by recombination under the assumption that the thermosphere is optically thin to the emitted photons. This implies that the ionization potential energy $I_s$ never contributes to heating at any levels. We also considered the influence of Lyman~$\alpha$ cooling by excited H, although this cooling rate is uncertain and likely to be low for HD209458b. We discuss the effect of different cooling rates further in Section~\ref{subsc:tempvel}. \subsubsection{Lower boundary conditions} \label{subsc:photochemistry} As stated above, we placed the lower boundary of the hydrodynamic model at $p_0 = $~1 $\mu$bar and did not include H$_2$ or other molecules in the model. This decision was motivated by photochemical calculations for HD209458b (Lavvas et al., \textit{in preparation}) that we used to constrain the lower boundary condition. The photochemical model was originally developed for the atmosphere of Titan \citep{lavvas08a,lavvas08b} but it was recently expanded to simulate EGP atmospheres. It calculates the chemical composition from the deep troposphere (1000 bar) up to the thermosphere above the 0.1 nbar level by solving the coupled continuity equations for all species based on a database of $\sim$1,500 reaction rate coefficients and 103 photolysis processes. Forward and reverse rates are included for each reaction with the ratio of the rate coefficients defined by thermochemical data. Thus, the results are consistent with thermochemical equilibrium at deep atmospheric levels but differences develop at higher altitudes due to photolysis, diffusion, and eddy mixing. At the lower boundary the chemical abundances of the main species (H, C, N, and O) are set to their thermodynamic equilibrium values and, depending on the vertical temperature profile and their abundances, species are allowed to condense. Figure~\ref{fig:photochemistry} shows the mixing ratios of H$_2$, H, H$_2$O, O, CH$_4$, CO, CO$_2$, and C as a function of altitude from the photochemical model. In general, the results are similar to those of \citet{moses11}. The H$_2$/H transition is located near 1 $\mu$bar. At lower pressures, the mixing ratio of H$_2$ decreases rapidly with altitude and falls below 0.1 above the 0.1~$\mu$bar level. In agreement with \citet{garciamunoz07}, the dissociation of H$_2$ is caused by dissociation of H$_2$O, which leads to the production of OH radicals that attack the H$_2$ molecules. We note that the exact location of the H$_2$/H transition depends on the temperature profile and, depending on the thermal structure, it could occur even below the 1 $\mu$bar level. \begin{figure} \centering \includegraphics[width=0.7\textwidth]{figure1.pdf} \caption{Mixing ratios of key oxygen and carbon-bearing species in the dayside atmosphere of HD209458b (Lavvas et al., \textit{in preparation}).} \label{fig:photochemistry} \end{figure} The major oxygen-bearing molecules, CO and H$_2$O, have roughly equal abundances from 10 to 10$^{-5}$ bar. This is in line with thermochemical equilibrium at the temperatures and pressures relevant to HD209458b \citep{lodders02}. H$_2$O and CO are effectively dissociated at pressures lower than 0.3 and 0.1 $\mu$bar, respectively. We note that molecular abundances could be significant at 0.1--1 $\mu$bar, and technically the results of the hydrodynamic calculations are only valid above the 0.1 $\mu$bar level. The presence of molecules such as H$_2$, H$_2$O, and CO can lead to enhanced UV heating as well as efficient radiative cooling by H$_3^+$, H$_2$O and CO in the 0.1--1 $\mu$bar region. The photochemical model also includes the chemistry of silicon. Due to condensation into forsterite and enstatite \citep[e.g][]{visscher10}, the abundance of Si in the observable atmospheres of giant planets was thought to be negligible and thus the photochemistry of silicon in planetary atmospheres has not been studied before. The photochemical calculations indicate that, in agreement with thermochemical calculations \citep{visscher10}, SiO is the dominant silicon-bearing gas. SiO dissociates in the thermosphere at a similar pressure level as H$_2$O and CO. We note that the detection of Si$^{2+}$ in the thermosphere implies that silicon does not condense in the atmosphere of HD209458b (Paper II). In addition to composition, lower boundary conditions are required for temperature and velocity. We specified $T_0$ and $p_0$ at the lower boundary, and used them to calculate $\rho_0$ from the ideal gas law. The steady state continuity equation $\rho_0 v_0 r^2 = F_c$, where $F_c$ is the flux constant, was used to calculate the velocity $v_0$ at the lower boundary during each time step. The flux constant is solved self-consistently by the model, and it depends largely on the terms in the energy equation. In general we assumed that $T_0 \approx$~1,300 K, which is close to the effective temperature of the planet. We note that this temperature is largely unconstrained. Radiative transfer models for close-in EGPs \citep[e.g.,][and references therein]{showman09} do not account for heating by stellar UV radiation or possible opacity sources created by photochemistry \citep[e.g.,][]{zahnle09}. Therefore these models may not accurately predict the temperature at the base of the thermosphere. \citet{sing08a,sing08b} used observations of the Na D line profile to constrain the temperature profile in the upper atmosphere. They suggested that Na condenses into clouds near the 3 mbar level, and thus predicted a deep minimum in temperature in this region that is required for condensation. The detection of Si$^{2+}$ implies that condensation of Na in the lower atmosphere is unlikely (Paper II), and thus this result is unreliable. Relying on the same observations, \citet{vidalmadjar11a,vidalmadjar11b} predicted that the temperature increases steeply from 1,300 K to 3,500 K near the 1 $\mu$bar level. However, their method to retrieve the temperature relies on the density scale height of Na to express the optical depth along the line of sight (LOS). This is not consistent with the argument that Na is depleted above the 3 mbar level. If such a depletion takes place, the density scale height of Na is not the same as the scale height of the atmosphere and it cannot be used to retrieve temperatures. \subsubsection{Upper boundary conditions} \label{subsc:upbound} Previous models of the thermosphere disagree on the details of the density and temperature profiles \citep[e.g.,][]{yelle04,tian05,garciamunoz07,murrayclay09}. This is partly due to different boundary conditions, although assumptions regarding the heating rates and composition are probably more important (see Section~\ref{subsc:tempvel}). Unfortunately, the planetary wind equations can have an infinite number of both subsonic and supersonic solutions. In time-dependent models, the upper boundary conditions in particular can determine if the solution is subsonic or supersonic, and they can alter the temperature and velocity profiles \citep[e.g.,][]{garciamunoz07}. The choice of proper boundary conditions is therefore important. \citet{volkov11a,volkov11b} studied the escape of neutral atmospheres under different circumstances by using the kinetic Monte Carlo (DSMC) method. Because the fluid equations are a simplification of the kinetic equations at low values of $Kn$, the hydrodynamic model should ideally be consistent with the DSMC results both above and below the exobase. \citet{volkov11a,volkov11b} found that the nature of the solutions depends on the thermal escape parameter $X_0 = G M_p m / k T_0 r_0$ and the Knudsen number $Kn_0$ at the lower boundary $r_0$ of a region where diabatic heating is negligible. They argued that hydrodynamic escape is possible when $X_0 <$~2--3 \citep[see also][]{opik63,hunten73}. When $X_0 \gtrsim$~3, on the other hand, the sonic point is at such a high altitude that the solution is practically subsonic and with $X_0 \gtrsim$~6 the escape rate is similar to the thermal Jeans escape rate. The results of the DSMC calculations can be incorporated into hydrodynamic models with appropriate upper boundary conditions. \citet{volkov11a,volkov11b} suggest that the modified Jeans escape rate, which is based on the drifting Maxwellian velocity distribution function, is a good approximation of the DSMC results in fluid models, consistent with \citet{yelle04}. In order to contrast the modified Jeans upper boundary conditions (hereafter, the modified Jeans conditions) with other possibilities, we used them and the extrapolated upper boundary conditions (hereafter, the `outflow' conditions) adopted by \citet{tian05} and \citet{garciamunoz07} in our simulations. In general, we placed the upper boundary at 16 R$_p$. The impact of the boundary conditions is discussed in Section~\ref{subsc:boundaries}. We formulated the outflow conditions simply by extrapolating values for density, temperature and velocity with a constant slope from below. For the modified Jeans conditions, we calculated the effusion velocity $v_s$ at the upper boundary separately for each species by using equation (9) from \citet{volkov11b}. Using this equation violates the conservation of electric charge at the upper boundary because the small mass of the electrons causes their velocity to be much larger than the velocity of the protons. In reality charge separation is prevented by the generation of an ambipolar electric field that ensures that the vertical current is zero at the upper boundary. This electric field causes the ions to escape faster while it slows the electrons down. Effectively this lowers the escape velocity ($v_{\text{esc}} = \sqrt{2 GM/r}$) of the ions and increases the escape velocity of the electrons. In order to incorporate the ambipolar electric field in the modified Jeans conditions we expressed the Jeans parameters for ions and electrons as: \begin{eqnarray} X_i&=&\frac{GMm_i}{kTr} - \frac{\phi_e q_i}{k T} \\ X_e&=&\frac{GMm_e}{kTr} + \frac{\phi_e |q_e|}{k T} \end{eqnarray} where $\phi_e$ is the ambipolar electric potential, $q_{i,e}$ is the electric charge and subscripts $i$ and $e$ stand for electrons and ions, respectively. We used these Jeans parameters to calculate the effusion velocities for the electrons and ions, and then solved iteratively for $\phi_e$ by using the condition of zero current i.e., $n_e |q_e| v_e = \sum_i n_i q_i v_i$. This approach is consistent with kinetic models for the solar and polar winds \citep{lemaire71a,lemaire71b}. Having obtained the correct effusion velocities for the charged and neutral species, we evaluated the mass weighted outflow velocity from: \begin{equation} v = \frac{1}{\rho} \sum_s \rho_s v_s \end{equation} and used this velocity as an upper boundary condition. The values of temperature and density that are required for this calculation were extrapolated from below. As the model approaches steady state, the solution approaches a value of $v$ at the upper boundary that is consistent with the modified Jeans velocity. \subsubsection{Numerical methods} We solved the equations of motion on a grid of 400--550 levels with increasing altitude spacing. The radius $r_n$ from the center of the planet at level $n$ is thus given by a geometric series \citep[e.g.,][]{garciamunoz07}: \begin{equation} r_n = r_1 + \sum_{i=1}^{n-1} f_c^i \delta z_0 \end{equation} where $r_1 =$~1.08 R$_p$, $\delta z_0 =$~10 km, and $f_c =$~1.014. We solved the equations of motions in two parts, separating advection (Eulerian terms) from the other (Lagrangian) terms. The Lagrangian solution is performed first, and all variables are updated before advection. We used the flux conservative van Leer scheme \citep[e.g.,][]{vanleer79} for advection, and the semi-implicit Crank-Nicholson scheme \citep[e.g,][]{jacobson99} to solve for viscosity and conduction in the momentum and energy equations, respectively. We employed a time step of 1 s in all of our calculations. Despite the sophisticated numerical apparatus, the model is still occasionally unstable, particularly in the early stages of new simulations. The primary source of the instabilities are pressure fluctuations (sound waves) that are not balanced by gravity. We used a two-step Shapiro filter \citep{shapiro70} periodically to remove numerical instabilities. We assumed that a steady state has been reached once the flux constant $F_c$ is constant with altitude and the flux of energy is approximately conserved. \section{Results} \label{sc:results} \subsection{Temperature and velocity profiles} \label{subsc:tempvel} In this section we constrain the range of mean temperatures and velocities based on stellar heating in the thermosphere of HD209458b and similar close-in EGPs. We discuss the general dependency of the temperature and velocity profiles on the net heating efficiency and stellar flux, and relate this discussion to new temperature and velocity profiles for HD209458b that are based on realistic photoelectron heating efficiencies calculated specifically for close-in EGPs. This discussion is necessary because the temperature and velocity profiles from previous models of the upper atmosphere differ significantly, and the differences affect the density profiles of the observed species (see Section~\ref{subsc:denprofs}). As an example, Figure~\ref{fig:modelcomp} shows the temperature profiles based on several earlier models. In addition to boundary conditions, the discrepancies evident in this figure arise from different assumptions about the heating rates. \begin{figure} \centering \includegraphics[width=0.7\textwidth]{figure2.pdf} \caption{Some examples of temperature profiles from earlier models of the upper atmosphere of HD209458b. The solid line is from Figure 1 of \citet{yelle04}, the dotted line is from the C2 model in this work (see Section~\ref{subsc:cecchieff}), the dashed line is the atomic hydrogen model from Figure 11 of \citet{tian05}, the dashed-dotted line is the SP model from Figure 3 of \citet{garciamunoz07}, and the dashed-triple-dotted line is from Figure 1 of \citet{murrayclay09}}. \label{fig:modelcomp} \end{figure} \subsubsection{General dependency} \label{subsc:gendep} We note that the thermal structure in the upper atmospheres of the giant planets in the solar system is not very well understood despite modeling and observations that are far more extensive than those available for extrasolar planets \citep[e.g.,][]{yelle04b}. It is therefore useful test the reaction of the model to different heating rates and profiles. We used our model to calculate temperature and velocity profiles based on different heating efficiencies and stellar fluxes. These profiles are shown in Figure~\ref{fig:gentempvel}. First, we used the average solar spectrum (Section~\ref{subsc:hydromodel}) and varied the net heating efficiency $\eta_{\text{net}}$ from 0.1 to 1. Second, we multiplied the solar spectrum by factors of 2x, 10x, and 100x, and used $\eta_{\text{net}} =$~0.5. The range of enhanced fluxes covers solar maximum conditions and stars that are more active than the sun. In each case we assumed that $\eta_{\text{net}}$ is constant and does not vary with altitude. We used planetary parameters of HD209458b and set the upper boundary to 16 $R_p$ with outflow boundary conditions, and the lower boundary to the 1 $\mu$bar level with a temperature of 1,300 K. \begin{figure} \centering \includegraphics[width=0.7\textwidth]{figure3a.pdf} \includegraphics[width=0.7\textwidth]{figure3b.pdf} \caption{Temperature (a) and velocity (b) profiles in the upper atmosphere of HD209458b based on different heating efficiencies and stellar XUV fluxes. The solid lines show models based on the average solar flux with $\eta_{\text{net}}$ of 0.1, 0.3, 0.5, 0.8, and 1 (from bottom to top). The dashed lines show models with $\eta_{\text{net}} =$~0.5 and stellar flux of 2x, 10x, and 100x the solar average flux (in order of increasing peak temperature).} \label{fig:gentempvel} \end{figure} A net heating efficiency of 50 \% is similar to the heating efficiency in the Jovian thermosphere \citep{waite83}, and we may consider this as a representative case of a typical gas giant (hereafter, the H50 model). The maximum temperature in the H50 model is 11,500 K and the temperature peak is located near 1.5 $R_p$ ($p = 0.3$ nbar). As $\eta_{\text{net}}$ varies from 0.1 to 1, the peak shifts from 1.4 $R_p$ (0.5 nbar) to 1.9 $R_p$ (0.1 nbar) and the maximum temperature varies from 10,000 K to 13,200 K. It is interesting to note that the temperature profile depends strongly on the heating efficiency but the location of the peak and the maximum temperature depend only weakly on $\eta_{\text{net}}$. This is because the vertical velocity increases with heating efficiency, leading to more efficient cooling by faster expansion that controls the peak temperature while enhanced advection and high altitude heating flatten the temperature gradient above the peak. As a result, the temperature profile is almost isothermal when $\eta_{\text{net}} =$~1. It is also interesting that the temperature profiles in the models that are based on $\eta_{\text{net}} =$~0.5 and the solar flux enhanced by factors of 2--100 differ from models with the average solar flux and a higher heating efficiency. For instance, one might naively assume that a model with $\eta_{\text{net}} =$~0.5 and 2x the average solar flux would be similar to a model with the average solar flux and $\eta_{\text{eta}} =$~1. Suprisingly, this is not the case -- despite the fact that the mass loss rates from these models are identical. This is because of the way radiation penetrates into the atmosphere -- doubling the incoming flux is not the same as doubling the heating rate at each altitude for the same flux. As the stellar flux increases further, the temperature peak shifts first to higher altitudes, and then to lower altitudes so that for the 100x case the peak is located again near 1.5 $R_p$. Despite the hugely increased stellar flux, the peak temperature is only 18,300 K for the 100x case. This is again because the enhanced adiabatic and advective cooling driven by faster expansion control the temperature even in the absence of efficient radiative cooling mechanisms. \citet{koskinen10} suggested that the mean temperature of the thermosphere below 3 $R_p$ can be constrained by observations, and used their empirical model to fit temperatures to the H Lyman~$\alpha$ transit data \citep{vidalmadjar03,benjaffel07,benjaffel08}. A quantity that can be compared with their results is the pressure averaged temperature of the hydrodynamic model, which is given by: \begin{equation} \overline{T_p} = \frac{\int_{p_1}^{p_2} T( p ) \ \text{d} (\ln p)}{\ln (p_2/p_1)}. \label{eqn:meantemp} \end{equation} For $\eta_{\text{net}}$ ranging from 0.1 to 1, the pressure averaged temperature below 3 $R_p$ varies from 6,200 K to 7,800 K for the average solar flux. In the H50 model the pressure averaged temperature is 7,000 K. We note that $\overline{T}_p$ is a fairly stable feature of our solutions -- in contrast to the details of the temperature profile and velocities it is relatively insensitive to different assumptions about the boundary conditions and heating efficiencies. Obviously, $\overline{T}_p$ depends on the stellar flux, although it only increases to 9,800 K in the 100x case. \citet{koskinen10} inferred a mean temperature of 8,250 K in the thermosphere of HD209458b with $p_0 =$~1 $\mu$bar (the M7 model). Taken together with our results based on solar XUV fluxes, this implies a relatively high heating efficiency. Alternatively, with $\eta_{\text{net}} =$~0.5 it may imply that the XUV flux of HD209458b is 5--10 times higher than the corresponding solar flux. This type of an enhancement is not impossible. The activity level of the star depends on its rotation rate, and the rotation rate of HD209458 may be twice as fast as the rotation rate of the sun \citep{silvavalio08}. However, the uncertainty of the observed H Lyman~$\alpha$ transit depths accommodates a range of temperatures, and thus we are unable to derive firm constraints on the heating rates from the observations. In general, though, the pressure averaged temperature provides a useful connection between observations and temperatures predicted by models that can be exploited to constrain heating rates. The effect of changing the heating efficiency on the velocity profile is quite dramatic. As $\eta_{\text{net}}$ ranges from 0.1 to 1 (with the average solar flux), the velocity at the upper boundary increases from 2.6 km~s$^{-1}$ to 25 km~s$^{-1}$. However, the velocity does not increase linearly with stellar flux or without a bound -- in the 100x case the velocity at the upper boundary is only 30 km~s$^{-1}$. An interesting qualitative feature of the solutions is that the sonic point moves to a lower altitude with increasing heating efficiency or stellar flux. With $\eta_{\text{net}} =$~0.1 the isothermal sonic point is located above the upper boundary whereas with $\eta_{\text{net}} =$~1 it is located at 4 $R_p$. This behavior is related to the temperature gradient and it is discussed further in Section~\ref{subsc:sonicpoint}. Basically the sonic point, when it exists, moves further from the planet as the high altitude heating rate decreases. It is now clear that assumptions regarding the heating efficiency and radiative transfer have a large impact on the temperature and velocity profiles and the results from the previous models reflect this fact (see Figure~\ref{fig:modelcomp}). The differences between models have implications on the interpretation of observations. For instance, \citet{vidalmadjar03} and \citet{linsky10} suggested that the UV transit observations probe the velocity structure of the escaping gas. Obviously, the nature of this velocity structure depends on the properties of the upper atmosphere. On the other hand, \citet{benjaffel10} argued that the observations point to a presence of hot energetic atoms and ions within the Roche lobe of the planet. We believe that it is important to properly quantify the role of stellar heating in creating the hot, escaping material before other options are pursued. This means that detailed thermal structure calculations that rely on a proper description of photoelectron heating efficiencies are required. Below we discuss a new approach to modeling the temperature profile in the thermosphere of HD209458b and its impact on the velocity and density profiles. \subsubsection{Energy balance and temperatures on HD209458b} \label{subsc:cecchieff} In the previous section we discussed models where the net heating efficiency $\eta_{\text{net}}$ was fixed at a constant value at all altitudes. In this section we discuss more realistic models of HD209458b that rely on new approximations of photoelectron heating efficiency and derive an estimate of $\eta_{\text{net}}$ based on these models. Here we also include radiative cooling from recombination and, in one case, H Lyman~$\alpha$ emissions by excited H \citep{murrayclay09}. Our aim was to calculate the most likely range of temperatures in the thermosphere of HD209458b based on average solar fluxes. Figure~\ref{fig:tempvel} shows the temperature and velocity profiles at 1--5 $R_p$ based on different approximations (see Table~\ref{table:models} for the input parameters). Model C1 assumes a constant photoelectron heating efficiency of 93 \% at all altitudes and photoelectron energies. This heating efficiency is appropriate for photoelectrons created by 50 eV photons at an electron mixing ratio of $x_e =$~0.1 \citep{cecchi09}. Model C2 is otherwise similar to C1 but the heating efficiency varies with photoelectron energy and altitude (see below). Models C3 and C4 are also based on C1, but C3 includes the substellar tidal forces in the equations of motion \citep[e.g.,][]{garciamunoz07} and C4 includes Lyman $\alpha$ cooling. All of these models are based on the outflow boundary conditions for temperature, velocity, and density. \begin{table}[htbp] \centering \caption{Model input parameters and results} \begin{tabular}{@{} ccccc @{}} \toprule \cmidrule( r ){1-5} Model$^{\textrm{a}}$ & $r_{\infty}$ (R$_p$)$^{\textrm{b}}$ & $\eta_{\textrm{net}}$$^{\textrm{c,d}}$ & $\dot{M}$ (10$^7$ kg~s$^{-1}$) & $\overline{T}_{p}$ (K)$^{\textrm{e}}$ \\ \midrule C1 & 16 E & 0.56 C & 5.6 & 7250 \\ C2 & 16 E & 0.44 V & 4.0 & 7200 \\ C3 & 16 E,T & 0.57 C & 6.4 & 6450 \\ C4 & 16 E & 0.46 C & 4.5 & 7110 \\ C5 & 36 J & 0.56 C & 5.6 & 7290 \\ C6 & 16 J & 0.45 V & 3.9 & 7310 \\ \bottomrule \end{tabular} \caption*{\small{$^a$All models assume $p_0 =$~10$^{-6}$ bar and $T_0 =$~1,300 K. \\ $^b$E - Outflow conditions, J - Modified Jeans conditions, T - Substellar tide. \\ $^c$Net heating efficiency (see Section~\ref{subsc:hydromodel}) i.e., the ratio of the net heating flux at all wavelengths to the unattenuated stellar flux (0.45 W~m$^{-2}$) at wavelengths shorter than 912~\AA. \\ $^d$C - Constant photoelectron heating efficiency, V - Varying photoelectron heating efficiency (see text). \\ $^e$Pressure averaged temperature below 3 $R_p$.}} \label{table:models} \end{table} \citet{cecchi09} also estimated the heating efficiencies for photoelectrons released by photons of energy $E \gtrsim$~50 eV at different values of the electron mixing ratio $x_e$. We used their heating efficiencies for $x_e =$~0.1 in the C2 model. They parameterized their results in terms of the vertical column density $N_{\textrm{H}}$ of H. We fitted the heating efficiency as a function of $N_{\textrm{H}}$ for 50 eV photons with a regular transmission function, and modified this function accordingly for different cutoff altitudes and heating efficiencies of photons with different energies [see Figures 3 and 4 of \citet{cecchi09}]. We note that $x_e \approx$~0.1 near the temperature peak of our models and thus the results are appropriate for our purposes. However, they are only applicable to photons with $E \gtrsim$~50 eV. We used simple scaling to estimate the heating efficiencies for low energy photons with $E <$~50 eV. \begin{figure} \centering \includegraphics[width=0.7\textwidth]{figure4a.pdf} \includegraphics[width=0.7\textwidth]{figure4b.pdf} \caption{Temperature (a) and velocity (b) profiles in the upper atmosphere of HD209458b based on different models (see Table~\ref{table:models} for the input parameters).} \label{fig:tempvel} \end{figure} As $N_{\textrm{H}}$ increases, the heating efficiency for 50 eV photons saturates at 93 \%. We assumed that the saturation heating efficiency for low energy photons is also 93 \%. In reality, this heating efficiency may be closer to 100 \% but the difference is small. In order to estimate the altitude dependence of the heating efficiency, we note that the rate of energy deposition by Coulomb collisions between photoelectrons of energy $E_p$ and thermal electrons with a temperature $T$ can be estimated from: \begin{equation} - \frac{\textrm{d} F_{E}}{\textrm{d} r} = L(E_p, e) \Phi_{pe} n_e \ \ \ [\textrm{eV} \ \textrm{cm}^{-3} \ \textrm{s}^{-1}] \end{equation} where $F_{E}$ is the flux of energy, $\Phi_{pe}$ is the photoelectron flux (cm$^{-2}$~s$^{-1}$), $n_e$ is the density of thermal electrons (cm$^{-3}$) and \[ L(E_p, e) = \frac{3.37 \times 10^{-12}}{n_e^{0.03} E_p^{0.94}} \left( \frac{E_p - E_e}{E_p - 0.53 E_e} \right)^{2.36} \ \ \ [\textrm{eV} \ \textrm{cm}^2], \] with $E_e =$~8.618~$\times$~10$^{-5} T_e$ is the stopping power \citep{swartz71}\footnote{Due to a historical precedent, the units here are in cgs.}. Assuming that all of the energy is deposited by electrons that are thermalized within a path element $dr$, we can estimate the e-folding length scale for thermalization of photoelectrons with different energies as follows: \begin{equation} \Lambda_{pe} \approx \frac{E_p}{n_e L}. \end{equation} We calculated $\Lambda_{pe}$ for different photoelectrons based on the C1 model, and compared the result with the vertical length scale $H$ of the atmosphere. The latter is either the scale height or $R_p$, depending on which is shorter. When $\Lambda_{pe}/H \gtrsim$~0.005--0.01 we assumed that the heating efficiency decreases with altitude according to the transmission function for 50 eV photons. The limiting value was chosen to obtain a rough agreement with the results of \citet{cecchi09} for 50 eV photons, and it implies that the heating efficiency approaches zero when $\Lambda_{pe}/H \gtrsim$~0.1. We parameterized the result in terms of the column density of H based on the density profiles of the C1 model, and connected our results for low energy photoelectrons smoothly with the results of \citet{cecchi09} for photons with $E \gtrsim$~50 eV. We then generated the C2 model from the C1 model with the new heating efficiencies. Figure~\ref{fig:peffs} shows the resulting heating efficiencies for 20, 30, 48, and 100 eV photons. \begin{figure} \centering \includegraphics[width=0.7\textwidth]{figure5.pdf} \caption{Heating efficiencies for photons of different energies (see text).} \label{fig:peffs} \end{figure} Figure~\ref{fig:penetration} shows the volume heating rate due to EUV photons of different energies as a function of pressure based on the C2 model. The maximum temperature of 12,000 K is reached near 1.5 R$_p$ ($p =$~0.6 nbar). This region is heated mainly by EUV photons with wavelengths between 200 and 900~\AA~($E =$~14--62 eV). The saturation heating efficiency of 93 \% for these photons is higher than the corresponding heating efficiency in the Jovian thermosphere \citep{waite83}. This is because of strong ionization that leads to frequent Coulomb collisions between photoelectrons and thermal electrons. Radiation with wavelengths shorter than 300~\AA~($E >$~40 eV) or longer than 912~\AA~(13.6 eV) penetrates past the temperature peak to the lower atmosphere. The heating efficiency for photons with $E >$~25 eV approaches zero at high altitudes where heating is mostly due to low energy EUV photons. The net heating efficiency for the C2 model is $\eta_{\text{net}} =$~0.44 (Table~\ref{table:models}), which is close to the H50 model. The location of the peak and maximum temperature in the C2 model agree with the H50 model, but the temperature at higher altitudes in the C2 model decreases much more rapidly with altitude than in the H50 model. \begin{figure} \centering \includegraphics[width=0.7\textwidth]{figure6.pdf} \caption{Volume heating rate as a function of pressure in the C2 model due to the absorption of stellar XUV radiation between 1 and 1000~\AA~in 200~\AA~bins.} \label{fig:penetration} \end{figure} Figure~\ref{fig:c2energy} shows the terms in the energy equation based on the C2 model. In line with previous studies, stellar heating is mainly balanced by adiabatic cooling. Advection cools the atmosphere at low altitudes below the temperature peak, whereas at higher altitudes it acts as a heating mechanism. In fact, above 2 $R_p$ the adiabatic cooling rate is higher than the stellar heating rate because thermal energy is transported to high altitudes by advection from the temperature peak. The radiative cooling term that is centered near 1.3 $R_p$ arises from recombination following thermal ionization. Recombination following photoionization is included implicitly in the model and the rate is not included in the output. Conduction is not significant at any altitude in the model. We note that the rates displayed in Figure~\ref{fig:c2energy} balance to high accuracy, thus implying that the simulation has reached steady state. \begin{figure} \centering \includegraphics[width=0.7\textwidth]{figure7.pdf} \caption{Volume heating rate based on the C2 model (absolute values). Advection acts as a cooling mechanism below 1.5 $R_p$ and a heating mechanism above this level.} \label{fig:c2energy} \end{figure} The differences between the C1 and C2 models are subtle. The peak temperatures are similar, and the temperature profiles generally coincide below 3 $R_p$. Above this radius the temperature in the C2 model decreases more rapidly with altitude than in the C1 model and subsequently the sonic point moves to higher altitudes above the model domain. The results indicate that the assumption of a constant photoelectron heating efficiency is appropriate below 3 $R_p$ whereas at higher altitudes it changes the nature of the solution. This should not be confused with the assumption of a constant $\eta_{\text{net}}$, which leads to a different temperature profile when compared with either C1 and C2 (see Figure~\ref{fig:gentempvel}). In general, the maximum and mean temperatures in models C1--C4 are relatively similar. Thus we conclude that the mean temperature in the thermosphere of HD209458b is approximately 7,000 K and the maximum temperature is 10,000--12,000 K. The substellar tide is included in the C3 model. We included it mainly to compare our results with previous models \citep{garciamunoz07,penz08,murrayclay09}. The substellar tide is not a particularly good representation of the stellar tide in a globally averaged sense. In reality, including tides in the models is much more complicated than simply considering the substellar tide \citep[e.g.,][]{trammell11}. Compared to the C1 model, the maximum temperature in the C3 model is cooler by $\sim$1,000 K and at high altitudes the C3 model is cooler by 1,000--2,000 K. The velocity is faster and hence adiabatic cooling is also more efficient. The substellar tide drives supersonic escape \citep[see also,][]{penz08} and the sonic point in the C3 model is at a much lower altitude than in the C1 model (see Section~\ref{subsc:sonicpoint}). However, it is not clear how the sonic point behaves as a function of latitude and longitude. Given that the tide is also likely to induce horizontal flows, it cannot be included accurately in 1D models. \citet{murrayclay09} argued that radiative cooling due to the emission of Lyman~$\alpha$ photons by excited H is important on close-in EGPs. The photons are emitted when the 2p level of H, which is populated by collisions with thermal electrons and other species, decays radiatively. We included this cooling mechanism in the C4 model by using the rate coefficient from \citet{glover07} that includes a temperature-dependent correction to the rate coefficient given by \citet{black81}. We also included an additional correction factor of 0.1 based on detailed level population and radiative transfer calculations by \citet{menager11}. The effect of Lyman $\alpha$ cooling is largest near the temperature peak where the C4 model is 1,500 K cooler than the C1 model, but generally the difference is not large. We note that the H Lyman $\alpha$ cooling rate here cannot be generalized as such to other EGPs because the level populations and opacities depend on the thermal structure and composition of the atmosphere. \subsubsection{Critical points} \label{subsc:sonicpoint} As we have pointed out, the location of the sonic point depends on the energy equation through the temperature profile. Here we show that the use of the isothermal approximation in estimating the location of the sonic point can lead to significant errors unless the atmosphere really is isothermal. The inviscid continuity and momentum equations can be combined to give an expression for the critical point $\xi_c$ of a steady-state solution \citep{parker65}: \begin{equation} -\frac{\textrm{d}}{\textrm{d} \xi} \left( \frac{c^2}{\xi^2} \right) = -\frac{1}{\xi^2} \frac{\textrm{d} c^2}{\textrm{d} \xi} + \frac{2 c^2}{\xi^3} = \frac{W^2}{\xi^4} \label{eqn:parkerpoint} \end{equation} where $\xi = r/r_0$, $c = \sqrt{k T/m}$ is the isothermal speed of sound, $W = GM_p/r_0$, and $m$ is the mean atomic weight. It is often assumed that the vertical velocity at the critical point is given by $v^2 = c^2 (\xi_c)$ so that the critical point coincides with the isothermal sonic point \citep{parker58}. However, \citet{parker65} suggested that subsonic solutions are also possible if the density at the base of the flow exceeds a critical value determined from the energy equation. In fact, he argued that conduction at the base of the corona may not be sufficient to drive a supersonic solar wind. This led him to suggest that supersonic expansion is possible only if there is significant heating of the corona over large distances above the base. For an isothermal atmosphere with a temperature $T_0$, equation (\ref{eqn:parkerpoint}) reduces to the famous result for the altitude of the sonic point \citep{parker58}: \begin{equation} \xi_c = \frac{W^2}{2 c_0^2} \label{eqn:parkerpoint2} \end{equation} where $W^2/c_0^2$ is the thermal escape parameter $X_0$ at the lower boundary $r_0$. The isothermal sonic point in the C1 model is located at 7.2 R$_p$ where $c(\xi_c) =$~7.2 km~s$^{-1}$. The volume averaged temperature of the C1 model below this point is approximately 7,100 K. Assuming that $r_0 =$~$R_p$, $T_0 =$~7,100 K, and $m = m_{\textrm{H}}$, $X_0 \sim$~16 and equation~(\ref{eqn:parkerpoint2}) yields $\xi_c \sim$~8. In this case the analytic result agrees fairly well with the hydrodynamic model if one accounts for partial ionization of the atmosphere by assuming that the mean atomic weight\footnote{The mean atomic weight can be less than 1 because electrons contribute to the number density but not significantly to the mass density.} is $m =$~0.9~$m_H$. On the other hand, the isothermal sonic point in the C2 model is at 15.4 $R_p$ where $c(\xi_c) =$~4.1 km~s$^{-1}$. This is because the temperature gradient of the model is steeper than the corresponding gradient in the C1 model. The volume averaged mean temperature below 15 $R_p$ in the C2 model is 3,900 K. With this temperature and $m = m_H$, equation~(\ref{eqn:parkerpoint2}) predicts a sonic point at 14.6 $R_p$. However, at 15 $R_p$ the atmosphere is mostly ionized and $m =$~0.6~$m_H$. With this value, the sonic point from equation~(\ref{eqn:parkerpoint2}) would be located at 8.8 $R_p$. These examples show that there are significant caveats to using equation~(\ref{eqn:parkerpoint2}) to estimate the altitude of the sonic point on close-in EGPs without accurate knowledge of the temperature and density profiles. A variety of outcomes are possible and it is difficult to develop a consistent criteria for choosing values of $T$ and $m$ that would produce satisfactory results. Another problem is that the atmosphere is not isothermal. In fact, the temperature gradient above the heating peak in models C1--C4 (Table~\ref{table:models}) is relatively steep, and in some cases it approaches the static adiabatic gradient ($T \propto r^{-1}$) as defined by \citet{chamberlain61}. Assuming that the temperature profile can be fitted with $c^2 = c_0^2/\xi^{\beta}$ above the heating peak, the estimated values of $\beta$ for the C1 and C2 models are 0.7 and 0.9, respectively. We note that the velocity in the C1 model exceeds the isentropic speed of sound ($c_{\gamma} = \sqrt{\gamma k T/m}$ where $\gamma =$~5/3) at 9.8 $R_p$ where $c_{\gamma} =$~8.7 km~s$^{-1}$. This altitude is significantly higher than the altitude of the isothermal sonic point. The velocity in the C2 model does not exceed the isentropic speed of sound below the upper boundary of 16 $R_p$. Thus the temperature profile has a significant impact on the nature of the solution and the escape mechanism. This means that estimating the altitude of the sonic point without observations and detailed models for guidance is almost certain to produce misleading results. Past models for the upper atmosphere of HD209458b have predicted a variety of altitudes for the sonic point. On the other hand, \citet{yelle04} pointed out that stellar heating in the thermosphere is mostly balanced by adiabatic cooling and our calculations confirm this. \citet{parker65} argued that the critical point stretches to infinity when $\beta \rightarrow$~1 i.e., as the temperature gradient is close to adiabatic. Based on this, we should perhaps expect that the sonic point on close-in EGPs is located at a fairly high altitude. This is confirmed by our hydrodynamic simulations. In all of our models except for one, the sonic point is located significantly above 5 $R_p$. The exception is the C3 model, which includes the substellar tide. The isentropic sonic point in this model is located at 3.9 $R_p$ where $c_{\gamma} =$~8.2 km~s$^{-1}$. This is because the substellar tide leads to a lower effective value of the potential $W$. However, the tidal potential depends on latitude and longitude, and the substellar results cannot be generalized globally. \subsubsection{Mass loss rates} \label{subsc:massloss} Here we evaluate the mass loss rates based on our models. We define the mass loss rate simply as: \begin{equation} \dot{M} = 4 \pi r^2 \rho v. \end{equation} We note that the solar spectrum that we used in this study contains the total flux of 4~$\times$~10$^{-3}$ W~m$^{-2}$ at wavelengths shorter then 912~\AA~(the ionization limit of H) when normalized to 1 AU. This value is close to the average solar flux of 3.9~$\times$~10$^{-3}$ W~m$^{-2}$ at the same wavelengths \citep[e.g.,][]{ribas05}. In order to simulate a global average, we divided the flux by a factor of 4 in the model. This means that the incident flux on HD209458b at 0.047 AU with wavelengths shorter than 912 \AA~in our model is 0.45 W~m$^{-2}$. The net heating efficiencies given in Table~\ref{table:models} are based on this value. Considering first the models with constant $\eta_{\text{net}}$ ranging from 0.1 to 1 (see Section~\ref{subsc:gendep}), the mass loss rate varies almost linearly with $\eta_{\text{net}}$ from 10$^7$ kg~s$^{-1}$ and 10$^8$ kg~s$^{-1}$ while the pressure averaged temperature below 3 $R_p$ changes only by 1,500 K. This is because in a hydrodynamic model such as ours the net energy has nowhere else to go but adiabatic expansion and cooling, and thus escape is energy-limited. The bulk of the energy is absorbed below 3 $R_p$, and the mass loss rate is largely set by radiative transfer in this region. The mass loss rate for HD209458b predicted by the C2 model is 4.1~$\times$~10$^7$ kg~s$^{-1}$ ($\eta_{\text{net}} =$~0.44). The C3 model has the highest mass loss rate, although this rate is only higher by a factor of 1.13 than the mass loss rate in the C1 model. Thus we predict a mass loss rate of 4--6~$\times$~10$^{7}$ kg~s$^{-1}$ from HD209458b based on the average solar flux at 0.047 AU. \citet{garciamunoz07} demonstrated that the mass loss rate is insensitive to the upper boundary conditions even when they have a large impact on the temperature and velocity profiles, particularly at high altitudes. Indeed, complex hydrodynamic models are not required to calculate mass loss rates under energy-limited escape as long as reasonable estimates of the net heating efficiency are available. It is also important to note that the current estimates of mass loss rates based on the observations \citep[e.g.,][]{vidalmadjar03} are not direct measurements. Instead, they are all based on different models. However, models that predict the same mass loss rate can predict different transit depths and models predicting different mass loss rates can match the observations equally well. Thus the models should not be judged on how well they agree with published mass loss rates but rather on how well they agree with the observed density profiles or transit depths. Hydrodynamic models with realistic heating rates are required to match the observations, and the mass loss rate then follows. The globally averaged mass loss rate of about 4--6~$\times$~10$^7$ kg~s$^{-1}$ from HD209458b agrees well with similar estimates calculated by \citet{yelle04,yelle06} and \citet{garciamunoz07}, respectively, but it is significantly larger than the value calculated by \citet{murrayclay09}. These authors report a mass loss rate of 3.3~$\times$~10$^7$ kg~s$^{-1}$ based on the substellar atmosphere. When multiplied by 1/4 this corresponds to a global average rate of about 8.3~$\times$~10$^6$ kg~s$^{-1}$. However, the substellar mass loss rate is also enhanced by tides, so a comparable global average taking this into account would be even less than 8.3~$\times$~10$^6$ kg~s$^{-1}$, which is already roughly a factor of 6 smaller than our calculations. The \citet{murrayclay09} models differ in many respects from the models described here including the treatment of boundary conditions and radiative cooling, the numerical approach, and the adoption of a gray approximation for stellar energy deposition. In order to explore the reason for the disagreement in escape rates, we have modified our model to implement the gray assumption by using the approach described in \citet{murrayclay09} (see Section~\ref{subsc:denprofs}). Specifically, we adopted an incident flux\footnote{By chance the incident flux is equal to the mean solar flux divided by 4 that we used as a `globally averaged' value. Here, however, it is taken to be the substellar value.} of 0.45 W~m$^{-2}$ and a mean photon energy of 20 eV. The mass loss rate based on the substellar atmosphere for this model is 2.8~$\times$~10$^7$ kg~s$^{-1}$, in good agreement with the \citet{murrayclay09} results. Thus, the difference between the \citet{murrayclay09} models and the others is due to the gray assumption, and the fact that they estimated the incident flux on HD209458b based on the solar flux integrated between 13 eV and 40 eV. This energy range contains only about 25 \% of the total solar flux at energies higher than 13.6 eV. Although not discussed by \citet{murrayclay09}, the restricted energy range is likely an attempt to account for the fact that the absorption cross section decreases with energy implying that photons of sufficiently high energy will be absorbed too deep in the atmosphere to affect escape or the thermal structure, or composition of the thermosphere. Whether this is true, however, depends on the composition and temperature of the atmosphere. The gray assumption also fails to include the fact that the net heating efficiency increases with higher photon energy. These difficulties highlight the basic problem with a gray model, that the results can only be accepted with confidence if verified by a more sophisticated calculation or direct observations. \subsubsection{Constraints from kinetic theory} \label{subsc:boundaries} Hydrodynamic models should be consistent with kinetic theory of rarefied media even if the modeled region is below the exobase. This is because the atmosphere is escaping to space, and the density decreases with altitude, falling below the fluid regime at some altitude above the exobase. Therefore the conditions in the exosphere affect the flow solutions even below the exobase. Inappropriate use of the hydrodynamic equations can lead one to overestimate the flow velocity and mass loss rate, and these errors also affect the temperature and density profiles. Thus it is important to demonstrate that the hydrodynamic solutions agree with constraints from kinetic theory \citep[e.g.,][]{volkov11a,volkov11b}. As an example, we calculated $Kn_0$ and $X_0$ (see Section~\ref{subsc:upbound}) based on the C1 and C2 models. The Knudsen number $Kn$ depends on the mean collision frequency, and it is much smaller than unity at all altitudes below 16 $R_p$. Thus the exobase is located above the model domain \citep[see also][]{murrayclay09}. Calculating values for $X_0$ is complicated by the broad stellar heating profile. We consider the region where stellar heating is negligible to be where the flux of energy \begin{equation} E_{\infty} = F_c \left( c_p T + \frac{1}{2} v^2 - \frac{G M_p}{r} \right) - \kappa r^2 \frac{\partial T}{\partial r} \end{equation} is approximately constant. This criteria is consistent with the equations of motion, and it means that $r_0$ that should be used to calculate $X_0$ is above the upper boundary of our model because significant stellar heating persists at all altitudes. Thus we evaluated values of $X$ near the upper boundary for guidance. We also calculated the values with both the mass of the proton ($X_{\textrm{H}}$) and the mean atomic weight ($X$). In the C1 model, $X_{\textrm{H}}$ decreases with altitude, and above 11.4 $R_p$ it has values of less than 3. The mean atomic weight near the upper boundary is $\sim$0.6 amu, and thus the general value of $X <$~2 above 11.1 $R_p$. The sonic point in the C1 model is below 11 $R_p$, and it is in a region where stellar heating is significant. In the C2 model, both $X$ and $X_{\text{H}}$ are greater than 3 at all altitudes below 16 $R_p$. In fact, $X$ increases with altitude above 9 $R_p$ because the temperature gradient parameter exceeds unity. Thus the values $X$ in the C1 and C2 models are consistent with the difference in altitude between the sonic points in these models (see Section~\ref{subsc:sonicpoint}). Indeed, our results show, in line with Parker's original ideas about the solar wind, that supersonic escape is possible if there is significant heating of the atmosphere over large distances above the temperature peak. Such heating flattens the temperature gradient and brings the sonic point closer to the planet. We note that there are some caveats to applying the simple criteria based on $Kn_0$ and $X_0$ to close-in extrasolar planets. The upper atmospheres of these planets are strongly ionized, and the DSMC simulations of \citet{volkov11a,volkov11b} apply only to neutral atmospheres. Partly due to ionization, the collision frequencies in the thermospheres of close-in planets are also high. Further, the atmospheres are affected by a broad stellar heating profile in altitude whereas the DSMC calculations do not include any diabatic heating. However, the results of \citet{volkov11a,volkov11b} also indicate that consistency with kinetic theory can be enforced approximately by applying the modified Jeans conditions to the hydrodynamic model at some altitude close to the exobase. This result is likely to be more general, and it applies to ionized atmospheres as long as ambipolar diffusion is taken into account (see Section~\ref{subsc:upbound}). We compared the temperature and velocity profiles from the C1 and C2 models with results from similar models C5 and C6 that use the modified Jeans conditions. Note again that our version of the modified Jeans conditions includes the polarization electrostatic field that is required in strongly ionized media. Figure~\ref{fig:modjeans} shows the temperature and velocity profiles from the models. There is no difference between the C5 model and the C1 model as long as the upper boundary of the C5 model is at a sufficiently high altitude. In this case we extended it to 36 $R_p$. When the upper boundary is placed at lower altitudes, the flow decelerates and the temperature increases near the upper boundary. A comparison between the C2 and C6 models provides an example of the difference that can arise when the modified Jeans boundary conditions are used significantly below the exobase. A better agreement is achieved if the upper boundary of the C6 model is placed at a slightly higher altitude. In summary, we have shown that the C1 and C2 models are both consistent with kinetic theory. \begin{figure} \centering \includegraphics[width=0.7\textwidth]{figure8a.pdf} \includegraphics[width=0.7\textwidth]{figure8b.pdf} \caption{Temperature (a) and velocity (b) profiles in the upper atmosphere of HD209458b based on models with extrapolated and modified Jeans upper boundary conditions (see Table~\ref{table:models} for the input parameters).} \label{fig:modjeans} \end{figure} We note that extending the models to 16 $R_p$ or higher is not necessarily justified because it ignores the complications arising from the possible influence of the stellar tide, the stellar wind, and interactions of the flow with the magnetic field of the planet or the star. We placed the upper boundary at a relatively high altitude to make sure that the boundary is well above the region of interest, but generally we do not consider our results to be accurate above 3--5 $R_p$. Instead, our results provide robust lower boundary conditions for multidimensional models of the escaping material outside the Roche lobe of the planet. Such models often cannot include detailed photochemical or thermal structure calculations. The results from the more complex models can then be used to constrain the upper boundary conditions in 1D models. This type of an iteration is a complex undertaking, and it will be pursued in future work. \subsection{Density profiles} \label{subsc:denprofs} In this section we provide a qualitative understanding of the density profiles and transition altitudes that affect the interpretation of the observations. Based on the gas giants in the solar system it might be expected that heavy species undergo diffusive separation in the thermosphere. If this were the case, the transit depths in the O I, C II, and Si III lines \citep{vidalmadjar04,linsky10} should not be significantly higher than the transit depth at visible wavelengths. It is therefore important to explain why diffusive separation does not take place in the thermosphere of HD209458b, and to clarify why H and O remain mostly neutral while C and Si are mostly ionized. Also, doubly ionized species such as Si$^{2+}$ are not common in planetary ionospheres, and their presence needs to be explained. In order to do this, we modeled the ionization and photochemistry of the relevant species, and prove that diffusive separation does not take place. In order to illustrate the results, Figure~\ref{fig:denprofs} shows the density profiles of H, H$^+$, He, He$^+$, O, O$^+$, C, C$^+$, Si, Si$^+$, and Si$^{2+}$ from the C2 model. The location of the H/H$^+$ transition obviously depends on photochemistry, but it also depends on the dynamics of escape. With a fixed pressure at the lower boundary, a faster velocity leads to a transition at a higher altitude. Thus the transition occurs near 3.1 $R_p$ in the C2 model whereas in the C1 and C3 models it occurs at 3.8 $R_p$ and 5 $R_p$, respectively. These results disagree with \citet{yelle04} and \citet{murrayclay09} who predicted a lower transition altitude, but they agree qualitatively with the solar composition model of \citet{garciamunoz07}. They also agree with the empirical constraints derived by \citet{koskinen10} from the observations. \begin{figure} \centering \includegraphics[width=0.7\textwidth]{figure9a.pdf} \includegraphics[width=0.7\textwidth]{figure9b.pdf} \caption{Density profiles in the upper atmosphere of HD209458b based on the C2 model (see Table~\ref{table:models} for the input parameters).} \label{fig:denprofs} \end{figure} Once again, the differences between the earlier models and our work arise from different boundary conditions, and assumptions regarding heating rates and photochemistry. We demonstrate this by reproducing the results of \citet{murrayclay09} with our model. In order to do so, we set the lower boundary to 30 nbar with a temperature of 1,000 K, and included the substellar tide in the equations of motion. We only included H, H$^+$, and electrons in the model, and used the recombination rate coefficient and Lyman~$\alpha$ cooling rate from \citet{murrayclay09}. We also calculated the heating and ionization rates with the gray approximation by assuming a single photon energy of 20 eV for the stellar flux of 0.45 W m$^{-2}$ at the orbital position of HD209458b. Figure~\ref{fig:mc09} shows the density profiles of H and H$^+$ based on this model (hereafter, the MC09 model). The H/H$^+$ transition in the MC09 model occurs near 1.4 $R_p$. If we replace the gray approximation with the full solar spectrum in this model, the H/H$^+$ transition moves higher to 2--3 $R_p$. This is because photons with different energies penetrate to different depths in the atmosphere, extending the heating profile in altitude around the heating peak. This is why the temperature at the 30 nbar level in the C2 model is 3,800 K and not 1,000 K. In order to test the effect of higher temperatures in the lower thermosphere, we extended the MC09 model to $p_0 =$~1 $\mu$bar (with $T_0 =$~1,300 K) and again used the full solar spectrum for heating and ionization. With these conditions, the H/H$^+$ transition moves up to 3.4 $R_p$, in agreement with the C2 model. We conclude that the unrealistic boundary conditions and the gray approximation adopted by \citet{murrayclay09} and \citet{guo11} lead to an underestimated overall density of H and an overestimated ion fraction. Thus their density profiles yield a H Lyman $\alpha$ transit depth of the order of 2--3 \% i.e., not significantly higher than the visible transit depth. We note that \citet{yelle04} also predicted a relatively low altitude of 1.7 $R_p$ for the H/H$^+$ transition -- despite the fact that his model does not rely on the gray approximation and the lower boundary is in the deep atmosphere. The reason for the low altitude of the H/H$^+$ transition in this case is the neglect of heavy elements. In the absence of heavy elements, H$_3^+$ forms near the base of the model and subsequent infrared cooling balances the EUV heating rates. This prevents the dissociation of H$_2$ below the 10 nbar level. In reality, reactions with OH and thermal decomposition dissociate H$_2$ near the 1 $\mu$bar level (see Section~\ref{subsc:photochemistry}) and cooling by H$_{3}^+$ is negligible at all altitudes. It should be noted that even if H$_2$ does not initially dissociate, H$_3^+$ can be removed from the lower thermosphere in reactions with carbon and oxygen species \citep[e.g.,][]{garciamunoz07} unless these species undergo diffusive separation. The subsequent lack of radiative cooling will then dissociate H$_2$ again near the 1 $\mu$bar level. \begin{figure} \centering \includegraphics[width=0.7\textwidth]{figure10.pdf} \caption{Density profiles of H and H$^+$ based on the MC09 model that is similar to that of \citet{murrayclay09} (see text). Compared with our models (see Figure~\ref{fig:denprofs}), the H/H$^+$ transition occurs at a significantly lower altitude. The difference arises from the lower boundary conditions and gray approximation to heating and ionization in the MC09 model.} \label{fig:mc09} \end{figure} In our models, charge exchange with oxygen (reactions R14 and R15 in Table~\ref{table:reactions}) dominates the photochemistry of H below 3 $R_p$ and charge exchange with silicon (R25, R26) is also important below 1.4 $R_p$. These reactions are secondary in a sense that they require the ions to be produced by some other mechanism. In fact, H$^+$ is mainly produced by photoionization (P1), although thermal ionization (R3) is also important near the temperature peak. The production rates are mainly balanced by loss to radiative recombination (R1). The net chemical loss timescale for H is longer than the timescale for advection above 1.7 $R_p$. This allows advection from below to replenish H at higher altitudes. The density profiles of O and O$^+$ are strongly coupled to H and H$^+$ by charge exchange \citep[see also][]{garciamunoz07}. As a result, the O/O$^+$ transition occurs generally near the H/H$^+$ transition. For instance, in the C2 model it is located near 3.4 $R_p$. The same is not true of carbon. The C/C$^+$ transition occurs at a much lower altitude than the H/H$^+$ and O/O$^+$ transitions. For instance, in the C2 model it is located near 1.2 $R_p$. C$^+$ is mainly produced by photoionization (P4), although thermal ionization (R8) and charge exchange with He$^+$ (R13) are also important near the temperature peak. The production is balanced by loss to radiative recombination (R10). The chemical loss timescale for C is shorter than the timescale for advection below 1.8 $R_p$. Thus advection is unable to move the C/C$^+$ transition to altitudes higher than 1.2 $R_p$. Silicon is almost fully ionized near the lower boundary of the model. Much of the Si$^{+}$ below 4 $R_p$ is produced by charge exchange of Si with H$^+$, He$^+$, and C$^+$ (R22, R23, R24). The low ionization potential of Si (8.2 eV) means that Si$^+$ can also be produced by thermal ionization (R18), and photoionization (P6) by stellar FUV radiation and X-rays that propagate past the EUV heating peak. Above 4 $R_p$, Si$^+$ is mostly produced by photoionization. \citet{linsky10} suggested that the balance of Si$^+$ and Si$^{2+}$ depends on charge exchange with H$^+$ and H, respectively, and our results confirm this. However, the location of the Si$^+$/Si$^{2+}$ transition also depends on the dynamics. For instance, in the C2 model it occurs near 5.8 $R_p$ while in the C1 model it occurs near 8.5 $R_p$. Thus slow outflow and high temperatures favor Si$^{2+}$ as the dominant silicon species as long as the flux constant is high enough to overcome diffusive separation (Paper II). We have now explained the presence of the atoms and ions that have been detected in the thermosphere of HD209458b. Due to advection and charge exchange, H and O are predominantly neutral up to about 3 $R_p$ and give rise to the observed transit depths in the H Lyman $\alpha$ and O I lines. Carbon, on the other hand, is ionized at a low altitude and thus C$^+$ is also detectable in the upper atmosphere. Si$^+$ is the dominant silicon species below 5 $R_p$, but charge exchange with H ensures that there is also a significant abundance of Si$^{2+}$ in the atmosphere. We note that these conclusions are only valid if the heavier species are carried along to high altitudes by the escaping hydrogen. We show that this is the case below in Section~\ref{subsc:ionescape}. \subsubsection{The EUV ionization peak (EIP) layer} \label{subsc:eip} \citet{koskinen10b} explored the properties of the ionospheres of EGPs at different orbital distances from a Sun-like host star by using a hydrostatic general circulation model (GCM) that also includes realistic heating rates, photochemistry, and transport of constituents. They predicted that the EIP layer on HD209458b is centered at 1.35 R$_p$ where the electron density is $n_e =$~6.4~$\times$~10$^{13}$ m$^{-3}$ and $x_e \sim$~3~$\times$~10$^{-2}$. In the C2 model, the EIP layer is centered at 1.3 R$_p$ ($p =$~2 nbar) where $n_e =$~4.4~$\times$~10$^{13}$ m$^{-3}$ and $x_e =$~3.7~$\times$~10$^{-2}$. The vertical outflow velocity at 1.3 R$_p$ is 90 m~s$^{-1}$. Thus the results of \citet{koskinen10b} were not significantly affected by the simplifying assumptions of the GCM. This means that hydrostatic GCMs can be extended to relatively low pressures as long as the escape rates are incorporated as boundary conditions. We also calculated the plasma frequency based on the electron densities in the C2 model. This constrains the propagation of possible radio emissions from the ionosphere. The ordinary plasma frequency $\omega_p / 2 \pi$ exceeds 12 MHz at all altitudes below 5 R$_p$ and reaches a maximum of almost 64 MHz in the EIP layer. This presents a limitation on the detection of radio emissions from the ionospheres of close-in EGPs. Any emissions that originate from the ionosphere at 1--5 R$_p$ and have frequencies lower than 10--70 MHz can be screened out by the ionosphere itself. We note that a detection of radio emissions from an EGP has not yet been achieved \citep[e.g.,][]{bastian00,lazio07,lecavelier11,griessmeier11}. Such a detection would be an important constraint on the magnetic field strength and the ionization state of the source region \citep[e.g.,][]{griesmeier07}. Models of the ionosphere are required to predict radio emissions from the possible targets. \subsubsection{The escape of heavy atoms and ions} \label{subsc:ionescape} In this section we verify \textit{a posteriori} that the velocity and temperature differences between different species in the thermosphere of HD209458b are small. This demonstrates that the single fluid approximation of the momentum and energy equations is valid, and that diffusive separation of the heavy species does not take place. Our model includes velocity differences between different species by including all of the relevant collisions between them through the inclusion of diffusive fluxes in the continuity equations. However, we have explicitly assumed that $T_n = T_i = T_e$, and this assumption in particular needs to be verified. Also, the diffusion approach to the continuity equation is only valid if the velocity differences between the species are reasonably small. We calculated the collision frequencies based on the C2 model, and found that collisions with neutral H dominate the transport of heavy neutral atoms such as O below 3.5 $R_p$. At altitudes higher than this, collisions with H$^+$ are more frequent. In Paper II we demonstrate that a mass loss rate of 6~$\times$~10$^6$ kg~s$^{-1}$ is required to prevent diffusive separation of O (the heaviest neutral species detected so far) in the thermosphere. The mass loss rate in our models is $\dot{M} >$~10$^7$ kg~s$^{-1}$ and thus O is dragged along to high altitudes by H. On the other hand, collisions with H$^+$ dominate the transport of heavy ions such as Si$^+$ as long as the ratio [H$^+$]/[H]~$\gtrsim$~10$^{-4}$ (Paper II). This explains why Coulomb collisions in our models are more frequent than heavy ion-H collisions at almost all altitudes apart from the immediate vicinity of the lower boundary. These collisions are much more efficient in preventing diffusive separation than collisions with neutral H. Figure~\ref{fig:difftimes} compares the timescale for diffusion $\tau_D$ for O and Si$^+$ with the timescale for advection $\tau_v$ based on the C2 model. In both cases, $\tau_D >> \tau_v$ and thus diffusion is not significant. This implies that there are no significant velocity differences between heavy atoms and hydrogen. We note that Coulomb collisions of doubly ionized species with H$^+$ are more frequent than collisions between a singly ionized species and H$^+$. Thus diffusion is even less significant for a species like Si$^{2+}$. We verified these results from our simulations by switching diffusion off in the model and rerunning the C2 model. As a result the density of the heavy atoms and ions increased slightly at high altitudes, but the differences are not significant -- the results were nearly identical to the density profiles of the original simulation. \begin{figure} \centering \includegraphics[width=0.7\textwidth]{figure11.pdf} \caption{Timescales for diffusion $\tau_D = H^2/D_{s}$ of O and Si$^+$, and for advection $\tau_v = H/v$ based on the C2 model. We calculated the diffusion coefficients in a mixture of H and H$^+$. The large scale height of the atmosphere and relatively high collision frequencies mean that diffusion is not significant ($\tau_D/\tau_v = v H/D_s >> 1$) in the thermosphere of HD209458b.} \label{fig:difftimes} \end{figure} We note that the atmosphere can also be mixed by vertical motions associated with circulation that are sometimes parameterized in one-dimensional models by the eddy diffusion coefficient $K_{zz}$ \citep[e.g.,][]{moses11}. This mechanism is efficient in bringing the heavy elements to the lower thermosphere but it is unlikely to mix the atmosphere up to 3 $R_p$ and beyond. Also, there is no generally accepted method of estimating the degree of global mixing based on circulation models, and most circulation models for EGPs do not adequately treat the relevant energy deposition and forcing mechanisms in the upper atmosphere. Thus there is considerable uncertainty over the values of $K_{zz}$ and rapid escape is a much more likely explanation for the lack of diffusive separation on HD209458b. In fact, the calculations of \citet{koskinen07} show that the temperature in the thermosphere of planets such as HD209458b is high enough to practically guarantee an effective escape rate. The only way to prevent this is to provide enough radiative cooling to offset most of the XUV flux, but there are no radiative cooling mechanisms efficient enough to achieve this in a thermosphere composed of atoms and ions. As we noted above, the temperatures of the electrons, ions, and neutrals are roughly equal in the thermosphere of HD209458b. In order to show this, we assumed that photoelectrons share their energy with thermal electrons, which then share this energy further with ions and neutrals. We also assumed that the collisions frequencies between the species are higher than the timescale for advection. If the velocity differences between the species are negligible, the steady state 5-moment energy equations for thermal electrons and ions \citep{schunk00} can be used to obtain the following approximations\footnote{Note that conduction and viscosity are not important in the thermosphere of HD209458b.}: \begin{eqnarray} T_e - T_i&\approx&\frac{1}{3} \frac{m_i}{m_e} \frac{q_R}{k n_e \nu_{ei}} \label{eqn:temp1} \\ T_i - T_n&\approx&\frac{1}{3} \frac{m_i + m_n}{m_i} \frac{q_R}{k n_i \nu_{in}} \label{eqn:temp2} \end{eqnarray} where $q_R$ is the volume heating rate, and $\nu_{ei}$ and $\nu_{in}$ are the electron-ion and ion-neutral momentum transfer collision frequencies, respectively. We calculated the temperature differences for H, H$^+$, and electrons based on the C2 model. The difference between the electron and ion temperatures decreases with altitude and is mostly less than 2 K. The difference between the ion and neutral temperatures, on the other hand, increases with altitude. The ion temperature is approximately 10 K higher than the neutral temperature near 5 $R_p$ and the difference reaches 150 K at 16 $R_p$. In both cases, the temperature differences are negligible compared to the temperature of the thermosphere. Further, the timescale for advection in the C2 model is always significantly longer than the relevant collision timescales. Thus we have shown that $T_e \approx T_i \approx T_n$ and that equations (\ref{eqn:temp1}) and (\ref{eqn:temp2}) are approximately valid. \section{Discussion and Conclusions} \label{sc:discussion} We have constructed a new model for the upper atmosphere of HD209458b in order to explain the detections of H, O, C$^+$, and Si$^{2+}$ at high altitudes around the planet \citep{vidalmadjar03,vidalmadjar04,linsky10}. There are many different interpretations of the observed transits in the H Lyman~$\alpha$ line \citep{vidalmadjar03,benjaffel07,benjaffel08,holstrom08,koskinen10}, and these interpretations rely on results from models of the upper atmosphere that are based on many uncertain assumptions \citep[see Section~\ref{subsc:gendep} and][for a review]{koskinen10}. Also, the detection of heavy atoms and ions in the thermosphere is not without controversy, and the detection of Si$^{2+}$ is particularly intriguing. Thus these observations present several interesting challenges to modelers. The observed transit depths are large, and substantial abundances of the relevant species are required to explain the observations. However, on every planet in the solar system heavier species are removed from the thermosphere by molecular diffusion and doubly ionized species are not commonly observed. Also, the observations imply that H and O remain mostly neutral in the thermosphere while C and Si are mostly ionized at a relatively low altitude. Hydrodynamic models coupled with chemistry and thermal structure calculations are required to explain the detection of these species in the upper atmosphere and the differences between their density profiles. Ours is the first such model that benefits from repeated detections of both neutral atoms and ions to constrain the composition and temperature. \citet{koskinen10} demonstrated that the H Lyman~$\alpha$ transit observations \citep{benjaffel07,benjaffel08} can be explained with absorption by H in the thermosphere if the base of the hot layer of H is near 1 $\mu$bar, the mean temperature within the layer is about 8,250 K, and the atmosphere is mostly ionized above 3 $R_p$. These parameters are based on fitting the data with a simple empirical model of the upper atmosphere. The density and temperature profiles from our new hydrodynamic model agree qualitatively with these constraints, demonstrating that the basic assumptions of \citet{koskinen10} are reasonable. This confirms once again that a comet-like tail \citep{vidalmadjar03} or energetic neutral atoms \citep{holstrom08} are not necessarily required to explain the H~Lyman~$\alpha$ observations. In line with recent results by \citet{moses11} and the empirical constraints mentioned above, we used a photochemical model of the lower atmosphere to show that H$_2$ dissociates near the 1 $\mu$bar level. Above this level, the lack of efficient radiative cooling and strong stellar EUV heating lead to high temperatures. We constrained the range of possible mean (pressure averaged) temperatures based on the average solar flux by using the hydrodynamic model to calculate temperatures with different heating efficiencies. For net heating efficiencies between 0.1 and 1, the mean temperature below 3 $R_p$ varies from 6,000 K to 8,000 K. This means that 8,000 K is a relatively strict upper limit on the mean temperature if the XUV flux of HD209458 is similar to the corresponding flux of the sun. A mean temperature of 8,250 K estimated from the observations implies the presence of an additional non-radiative heat source, or that the XUV flux from HD209458 is higher than the average solar flux. Given that our best estimate of the net heating efficiency is 0.44 (see Section~\ref{subsc:cecchieff}), the XUV flux of H209458 would have to be 5--10 times higher than the average solar flux to cause a mean temperature of 8,250 K (see Section~\ref{subsc:gendep}). If the mean XUV flux of HD209458 is generally higher than the solar flux and the observations took place during stellar maximum, such an enhancement is not impossible. This would also lead to higher outflow velocity and mass loss rate. However, the uncertainty in the observed transit depths is also large \citep{benjaffel08,benjaffel10}, and it can accommodate a range of temperatures. Therefore our reference model C2 with a mean temperature of 7,200 K also agrees qualitatively with the empirical constraints. In this respect, it is interesting to note that with 100x solar flux, the mean temperature is still only 9,800 K. Temperatures significantly higher than 8,000 K \citep[e.g.,][]{benjaffel10} therefore imply a strong non-radiative heat source. In contrast to the mean temperature, the velocity and details of the temperature profile depend strongly on the heating efficiency and stellar flux (see Section~\ref{subsc:gendep}). They are also sensitive to the upper and lower boundary conditions. This explains the large range of temperature and density profiles predicted by earlier models that arise from different boundary conditions and assumptions about the stellar flux, radiative transfer, and heating efficiencies. The differences highlight the need for accurate thermal structure calculations that are constrained by the available observations. These calculations are important because the density profiles of the detected species depend on the temperature and velocity profiles, and inappropriate assumptions made by the models can bias the interpretation of the observations. In the absence of stellar gravity, the location of the sonic point and the outflow speed also depend on the heating efficiency. As the heating efficiency increases from 0.1 (in models with the average solar flux), the high altitude temperature increases and the sonic point moves to lower altitudes, reaching down to 4 $R_p$ with a net heating efficiency of 1. We found that supersonic solutions are possible as long as there is significant heating over a large altitude range above the temperature peak. This conclusion is supported both by the hydrodynamic model and new constraints from kinetic theory \citep{volkov11a,volkov11b}. However, the isentropic sonic point of the C2 model is located above the model domain. In principle, this is an interesting result but it should be treated with caution. We used parameterized heating efficiencies for low energy photons, and the location of the sonic point is very sensitive to the temperature profile. Also, the stellar tide can enhance the escape rates at the substellar and antistellar points. We did not include this effect because it may produce horizontal flows that cannot be modeled in 1D. As long as the upper boundary is at a sufficiently high altitude, we found that the results based on the outflow boundary conditions and modified Jeans conditions are identical (see Section~\ref{subsc:boundaries}). This shows that our simulations are roughly consistent with kinetic theory. An agreement between these two types of boundary conditions on HD209458b is an interesting theoretical result. It shows that the boundary conditions for hydrodynamic escape are appropriate in this case. However, an upper boundary at 16 $R_p$ or higher is not necessarily justified for other reasons because we did not consider the effect of the possible planetary magnetic field, interaction of the atmosphere with the stellar wind, or horizontal transport \citep[e.g.,][]{stone09,trammell11}. We chose an upper boundary at a high altitude in order to preserve the integrity of the solution in our region of interest below 5 $R_p$. The purpose of this work is to model energy deposition and photochemistry in this region. These aspects are often simplified in more complex models to a degree that it may be difficult to separate the effect of multiple dimensions and other complications from differences arising simply because of different assumptions about heating efficiencies and chemistry. Also, the uncertainty in the observations does not necessarily justify the introduction of more free parameters to the problem until the basic properties of the thermosphere are better understood. However, technically we do not consider our solutions to be accurate far above 3--5 $R_p$. Instead, our results provide robust lower boundary conditions for more complex multidimensional models that characterize the atmosphere outside the Roche lobe of the planet. Results from such models can then be used to constrain the upper boundary conditions of the 1D models further. In order to model the density profiles of the detected species in the ionosphere, we assumed solar abundances of the heavy elements \citep{lodders03}, although this assumption can be adjusted as required to explain the observations (Paper II). As we already stated we found that H$_2$, H$_2$O, and CO dissociate above the 1 $\mu$bar level, releasing H, O, and C to the thermosphere \citep[see also][]{moses11}. We note that the detection of Si$^{2+}$ in the upper atmosphere implies that silicon does not condense into clouds of forsterite and enstatite in the lower atmosphere as argued by e.g., \citet{visscher10}. The dominant Si species is then SiO, which dissociates at a similar pressure level as the other molecules. In fact, practically all molecules dissociate below 0.1 $\mu$bar. This leads to an important simplification in hydrodynamic models of the thermosphere. The complex chemistry of molecular ions does not need to be included as long as the lower boundary is above the dissociation level. We found that the H/H$^+$ transition occurs near 3 $R_p$ or, depending on the velocity profile, at even higher altitudes. The O/O$^+$ transition is coupled to the H/H$^+$ transition through charge exchange reactions. Thus both H and O are mostly neutral up to the boundary of the Roche lobe at 3--5 $R_p$. In contrast, C is ionized near the 1 $\mu$bar level and C$^+$ is the dominant carbon species in the thermosphere. Si is also ionized near the 1 $\mu$bar level, and the balance between Si$^+$ and Si$^{2+}$ is determined by charge exchange with H$^+$ and H, respectively. Si$^+$ is the dominant silicon ion below 5 $R_p$ but the abundance of Si$^{2+}$ is also significant. We found that neutral heavy atoms are dragged to the thermosphere by the escaping H, while heavy ions are transported efficiently by the escaping H$^+$. Thus the advection timescale is much shorter than the diffusion timescale of the detected species, and diffusive separation does not take place in the thermosphere. We also verified that the neutral, ion, and electron temperatures are roughly equal. Taken together, these results imply that the thermospheres of close-in EGPs can differ fundamentally from the gas giant planets in the solar system. For instance, the thermosphere of HD209458b is composed mainly of atoms and atomic ions, and diffusive separation of the common heavy species is prevented by the escape of H and H$^+$. It is important to note, however, that results such as these cannot be freely generalized to other extrasolar planets. As in the solar system, each planet should be studied separately. For instance, the dissociation of molecules depends on the temperature profile that is shaped by the composition through radiative cooling and stellar heating. The mass loss rate and escape velocity, that determine whether diffusive separation takes place or not, depends on the escape mechanism that again depends on the temperature and composition of the upper atmosphere. The results from different models can only be verified by observations that are therefore required for multiple planets if we are to characterize escape in different systems and under different conditions.\\ We are grateful to A. Volkov for reading the manuscript and providing useful feedback. We thank H. Menager, M. Barthelemy, J.-M. Grie\ss meier, N. Lewis, D. S. Snowden, and C. Cecchi-Pestellini for useful discussions and correspondence. We also acknowledge the "Modeling atmospheric escape" workshop at the University of Virginia and the International Space Science Institute (ISSI) workshop organized by the team "Characterizing stellar and exoplanetary environments" for interesting discussions and an opportunity to present our work. The calculations for this paper relied on the High Performance Astrophysics Simulator (HiPAS) at the University of Arizona, and the University College London Legion High Performance Computing Facility, which is part of the DiRAC Facility jointly funded by STFC and the Large Facilities Capital Fund of BIS. SOLAR2000 Professional Grade V2.28 irradiances are provided by Space Environment Technologies.
\section{Introduction} \subsection{Motivation and setting} In analyzing multivariate time series data, collected in financial applications, monitoring of influenza outbreaks and other fields, it is often of key importance to accurately characterize dynamic changes over time in not only the mean of the different elements (e.g., assets, influenza levels at different locations) but also the covariance. As shown in Figure~\ref{f:intro}, it is typical in many domains to cycle irregularly between periods of rapid and slow change; most statistical models are insufficiently flexible to capture such locally varying smoothness in assuming a single bandwidth parameter. Inappropriately restricting the smoothness to be constant can have a major impact on the quality of inferences and predictions, with over-smoothing occurring during times of rapid change. This leads to an under-estimation of uncertainty during such volatile times and an inability to accurately predict risk of extremal events. \begin{figure}[t] \centering \includegraphics[height=9cm, width=16cm]{DAX.pdf} \put (-300,225) {DAX30: Squared log returns} \caption{\footnotesize{Squared Log-Returns of DAX30, using weekly data from $2004/07/19$, to $2012/06/25$.}} \label{f:intro} \end{figure} Let $Y_t = (Y_{t1},\ldots,Y_{tp})^T$ denote a random vector at time $t$, with $\mu(t) = \mbox{E}( Y_t )$ and $\Sigma(t) = \mbox{cov}( Y_t )$. Our focus is on Bayesian modeling and inference for the multivariate mean-covariance stochastic process, $\Gamma = \{ \mu(t),\Sigma(t), t \in \mathcal{T}\}$ with $\mathcal{T} \subset \Re_+$. Of particular interest is allowing locally-varying smoothness, meaning that the rate of change in the $\{ \mu(t),\Sigma(t) \}$ process is varying over time. To our knowledge, there is no previous proposed stochastic process for a coupled mean-covariance process, which allows locally-varying smoothness. A key to our construction is the use of latent processes, which have time-varying smoothness. This results in a {\em locally adaptive factor} (LAF) process. We review the relevant literature below and then describe our LAF formulation. \subsection{Relevant literature} There is a rich literature on modeling a $p \times 1$ time-varying mean vector $\mu(t)$, covering multivariate generalizations of autoregressive models (VAR, e.g. \citealp{Tsay:2005}), Kalman filtering \citep{Ka:1960}, nonparametric mean regression via Gaussian processes (GP) \citep{Ra:2006}, polynomial spline \citep{Hua:2002}, smoothing spline \citep{Ha:1990} and kernel smoothing methods \citep{Wo:2011}. Such approaches perform well for slowly-changing trajectories with constant bandwidth parameters regulating implicitly or explicitly global smoothness; however, our interest is allowing smoothness to vary locally in continuous time. Possible extensions for local adaptivity include free knot splines (MARS) \citep{Fri:1991}, which perform well in simulations but the different strategies proposed to select the number and the locations of knots (stepwise knot selection \citep{Fri:1991}, Bayesian knot selection \citep{Sm:1996} or via MCMC methods \citep{Geo:1993}) prove to be computationally intractable for moderately large $p$. Other flexible approaches include wavelet shrinkage \citep{Dono:1995}, local polynomial fitting via variable bandwidth \citep{Fan:1995} and linear combination of kernels with variable bandwidths \citep{Wo:2011}. There is a separate literature on estimating a time-varying covariance matrix $\Sigma(t)$. This is particular of interest in applications where volatilities and co-volatilities evolve through non constant paths. One popular approach estimates $\Sigma(t)$ via an exponentially weighted moving average (EWMA; see, e.g., \citealp{Tsay:2005}). This approach uses a single time-constant smoothing parameter $0<\lambda<1$, with extensions to accommodate locally-varying smoothness not straightforward due to the need to maintain positive semidefinite $\Sigma(t)$ at every time. To allow for higher flexibility in the dynamic of the covariances, generalizations of EWMA have been proposed including the diagonal vector ARCH model (DVEC), \citep{Boll:1988} and its variant, the BEKK model \citep{Eng:1995}. These models are computationally demanding and are not designed for moderate to large $p$. DCC-GARCH \citep{Eng:2002} improves the computational tractability of the previous approaches through a two-step formulation. However, the univariate GARCH assumed for the conditional variances of each time series and the higher level GARCH models with the same parameters regulating the evolution of the time varying conditional correlations, restrict the evolution of the variance and covariance matrices. PC-GARCH (\citealp{Din:1994} and \citealp{Burn:2005}) and O-GARCH \citep{Ale:2001} perform dimensionality reduction through a latent factor formulation (see also \citealp{Weide:2002}). However, time-constant factor loadings and uncorrelated latent factors constrain the evolution of $\Sigma(t)$. Such models fall far short of our goal of allowing $\Sigma(t)$ to be fully flexible with the dependence between $\Sigma(t)$ and $\Sigma(t+\Delta)$ varying with not just the time-lag $\Delta$ but also with time. In addition, these models do not handle missing data easily and tend to require long series for accurate estimation \citep{Burn:2005}. Accommodating changes in continuous time is important in many applications, and avoids having the model be critically dependent on the time scale, with inconsistent models obtained as time units are varied. \citet{Wi:2010} join machine learning and econometrics efforts by proposing a model for both mean and covariance regression in multivariate time series, improving previous work of \citet{Bru:1991} on Wishart processes in terms of computational tractability and scalability, allowing a more complex structure of dependence between $\Sigma(t)$ and $\Sigma(t+\Delta)$. Specifically, they propose a continuous time Generalised Wishart Process (GWP), which defines a collection of positive semi-definite random matrices $\Sigma(t)$ with Wishart marginals. Nonparametric mean regression for $\mu(t)$ is also considered via GP priors; however, the trajectories of means and covariances inherit the smooth behavior of the underlying Gaussian processes, limiting the flexibility of the approach in times exhibiting sharp changes. Even for iid observations from a multivariate normal model with a single time stationary covariance matrix, there are well known problems with Wishart priors motivating a rich literature on dimensionality reduction techniques based on factor and graphical models. There has been abundant recent interest in applying such approaches to dynamic settings. Refer to \citet{Naka:2012} and the references cited therein for recent literature on Bayesian dynamic factor models for multivariate stochastic volatility. Their approach allows the factor loadings to evolve dynamically over time, while including sparsity through a latent thresholding approach, leading to apparently improved performance in portfolio allocation. They utilize a time-varying discrete-time autoregressive model, which allows the dependence in the covariance matrices $\Sigma(t)$ and $\Sigma(t+\Delta)$ to vary as a function of both $t$ and $\Delta$. However, the result is an extremely richly parameterized and computationally challenging model, with selection of the number of factors proceeding by cross validation. Our emphasis is instead on developing continuous time stochastic processes for $\Sigma(t)$ and $\mu(t)$, which accommodate locally-varying smoothness. \citet{Fox:2011} propose an alternative Bayesian covariance regression (BCR) model, which defines the covariance matrix as a regularized quadratic function of time-varying loadings in a latent factor model, characterizing the latter as a sparse combination of a collection of unknown Gaussian process (GP) dictionary functions. Although their approach provides a continuous time and highly flexible model that accommodates missing data and scales to moderately large $p$, there are two limitations motivating this article. Firstly, their proposed covariance stochastic process assumes a stationary dependence structure, and hence tends to under-smooth during periods of stability and over-smooth during periods of sharp changes. Secondly, the well known computational problems with usual GP regression are inherited, leading to difficulties in scaling to long series and issues in mixing of MCMC algorithms for posterior computation. \subsection{Contribution and outline} Our proposed LAF process instead includes dictionary functions that are generated from nested Gaussian processes (nGP) \citep{Zhu:2012}. Such nGP reduces the GP computational burden involving matrix inversions from $O(T^3)$ to $O(T)$, with $T$ denoting the length of the time series, while also allowing flexible locally-varying smoothness. Marginalizing out the latent factors, we obtain a stochastic process that inherits these advantages. We also develop a different and more computationally efficient approach to computation under this new model and propose online implementation, which can accommodate streaming data. In Section 2, we describe LAF structure with particular attention to prior specification. Section 3 explores the main features of the Gibbs sampler for posterior computation and outlines the steps for a fast online updating approach. In Section 4 we compare our model to BCR and to some of the most quoted models for multivariate stochastic volatility, through simulation studies. Finally in Section 5 an application to stock market indices across countries is examined. \section{Locally Adaptive Factor Processes} \subsection{ Notation and motivation} Our focus is on defining a novel locally adaptive factor (LAF) process for $\Gamma = \{ \mu(t),\Sigma(t), t \in \mathcal{T}\}$. In particular, taking a Bayesian approach, we define a prior $\Gamma \sim P$, where $P$ is a probability measure over the space $\mathcal{P}$ of $p$-variate mean-covariance processes on $\mathcal{T}$. In particular, each element of $\mathcal{P}$ corresponds to a realization of the stochastic process $\Gamma$, and the measure $P$ assigns probabilities to a $\sigma$-algebra of subsets of $\mathcal{P}$. Although the proposed class of LAF processes can be used much more broadly, in conducting inferences in this article, we focus on the simple case in which data consist of vectors $y_i = (y_{i1},\ldots,y_{ip})^T$ collected at times $t_i$, for $i=1,\ldots,n$. These times can be unequally-spaced, or collected under an equally-spaced design with missing observations. An advantage of using a continuous-time process is that it is trivial to allow unequal spacing, missing data, and even observation times across which only a subset of the elements of $y_i$ are observed. We additionally make the simplifying assumption that \begin{eqnarray*} Y_{i}\sim \mbox{N}_{p}(\mu(t_{i}),\Sigma(t_{i})). \label{eq:0} \end{eqnarray*} It is straightforward to modify the methodology to accommodate substantially different observation models. \subsection{LAF specification} A common strategy in modeling of large $p$ matrices is to rely on a lower-dimensional factorization, with factor analysis providing one possible direction. Sparse Bayesian factor models have been particularly successful in challenging cases, while having advantages over frequentist competitors in incorporating a probabilistic characterization of uncertainty in the number of factors as well as the parameters in the loadings and residual covariance. For recent articles on Bayesian sparse factor analysis for a single large covariance matrix, refer to \citet{Bhat:2011}, \citet*{Pat:2012} and the references cited there-in. In our setting, we are instead interested in letting the mean vector and the covariance matrix vary flexibly over time. Extending the usual factor analysis framework to this setting, we say that $\Gamma = \{ \mu(t),\Sigma(t), t \in \mathcal{T}\} \sim \mbox{LAF}_{L,K}(\Theta, \Sigma_{0},\Sigma_{\xi} ,\Sigma_{A},\Sigma_{\psi}, \Sigma_{B})$ if \vspace{-5pt} \begin{subequations} \begin{align} \mu(t)&= \Theta \xi(t) \psi(t) \label{subeq1}\\ \Sigma(t)&= \Theta \xi(t) \xi(t)^T \Theta^T+\Sigma_{0} \label{subeq2} \end{align} \end{subequations} where $\Theta$ is a $p \times L$ matrix of constant coefficients, $\Sigma_{0}=\mbox{diag}(\sigma_{1}^{2},...,\sigma_{p}^{2})$, while $\xi(t)_{L \times K}$ and $\psi(t)_{K \times 1}$ are matrices comprising continuous dictionary functions evolving in time through nGP, $\xi_{lk}(t)\sim \mbox{nGP}([\Sigma_{\xi}]_{lk}=\sigma^2_{\xi_{lk}},[\Sigma_{A}]_{lk}=\sigma^2_{A_{lk}})$ and $\psi_{k }(t)\sim \mbox{nGP}([\Sigma_{\psi}]_{k}=\sigma^2_{\psi_{k}},[\Sigma_{B}]_{k}=\sigma^2_{B_{k}})$. Restricting our attention on the generic element $\xi_{lk}(t): \mathcal{T} \rightarrow \Re$ of the matrix $\xi(t)_{L\times K}$ (the same holds for $\psi_{k}(t):\mathcal{T} \rightarrow \Re$), the nGP provides a highly flexible stochastic process on the dictionary functions whose smoothness, explicitly modeled by their $m$th order derivatives $D^{m}\xi_{lk}(t)$ via stochastic differential equations (SDEs), is expected to be centered on a local instantaneous mean function $A_{lk}(t)$, which represents a higher-level Gaussian Process (GP), that induces adaptivity to locally-varying smoothing. Specifically, we let \vspace{-10pt} \begin{subequations} \begin{align} D^{m}\xi_{lk}(t)&=A_{lk}(t)+\sigma_{\xi_{lk}}W_{\xi_{lk}}(t), \quad m \in N, \quad m \geq 2, \label{subeq5}\\ D^{n}A_{lk}(t)&=\sigma_{A_{lk}}W_{A_{lk}}(t), \ \quad \quad \quad \quad \ n \in N, \quad \ n \geq1, \label{subeq6} \end{align} \end{subequations} where $\sigma_{\xi_{lk}} \in \Re^+$, $\sigma_{A_{lk}} \in \Re^+$, $W_{\xi_{lk}}(t):\mathcal{T} \rightarrow \Re$ and $W_{A_{lk}}(t):\mathcal{T} \rightarrow \Re$ are independent Gaussian white noise processes with mean $\mbox{E}[W_{\xi_{lk}}(t)]=\mbox{E}[W_{A_{lk}}(t)]=0$, for all $t \in \mathcal{T}$, and covariance function $\mbox{E}[W_{\xi_{lk}}(t)W_{\xi_{lk}}(t')]=\mbox{E}[W_{A_{lk}}(t)W_{A_{lk}}(t')]=1$ if $t=t'$, $0$ otherwise. This formulation naturally induces a stochastic process for $\xi_{lk}(t)$ with varying smoothness, where $\mbox{E}[D^{m}\xi_{lk}(t)|A_{lk}(t)]=A_{lk}(t)$, and initialization at $t_{1}$ based on the assumption \begin{eqnarray*} [\xi_{lk}(t_{1}),D^{1}\xi_{lk}(t_{1}),...,D^{m-1}\xi_{lk}(t_{1})]^{T} &\sim& \mbox{N}_{m}(0,\sigma^{2}_{\mu_{lk}}I_{m})\\ \ [A_{lk}(t_{1}),D^{1}A_{lk}(t_{1}),...,D^{n-1}A_{lk}(t_{1})]^{T} &\sim& \mbox{N}_{n}(0,\sigma^{2}_{\alpha_{lk}}I_{n})\end{eqnarray*} The Markovian property implied by SDEs in (\ref{subeq5}) and (\ref{subeq6}) represents a key advantage in terms of computational tractability as it allows a simple state space formulation. In particular, referring to \citet{Zhu:2012} for $m=2$ and $n=1$ (this can be easily extended for higher $m$ and $n$), and for $\delta_{i}=t_{i+1}-t_{i}$ sufficiently small, the process for $\xi_{lk}(t)$ along with its first order derivative $\xi_{lk}'(t)$ and the local instantaneous mean $A_{lk}(t)$ follow the approximated state equation \renewcommand{\arraystretch}{0.8} \begin{eqnarray} \left[ \begin{array}{c} \xi_{lk}(t_{i+1})\\ \xi'_{lk}(t_{i+1})\\ A_{lk}(t_{i+1}) \end{array} \right]= \left[ \begin{array}{ccc} 1&\delta_{i}&0\\ 0 &1&\delta_{i}\\ 0 &0&1 \end{array} \right] \left[ \begin{array}{c} \xi_{lk}(t_{i})\\ \xi'_{lk}(t_{i})\\ A_{lk}(t_{i}) \end{array} \right]+ \left[ \begin{array}{cc} 0&0\\ 1&0\\ 0&1 \end{array} \right] \left[ \begin{array}{c} \omega_{i,\xi_{lk}}\\ \omega_{i,A_{lk}}\\ \end{array} \right], \label{eq:11} \end{eqnarray} \renewcommand{\arraystretch}{0.5} where $[\omega_{i,\xi_{lk}}, \omega_{i,A_{lk}}]^T\sim \mbox{N}_{2}(0,V_{i,lk})$, with $V_{i,lk}=\mbox{diag}(\sigma^2_{\xi_{lk}} \delta_{i}, \sigma^2_{A_{lk}} \delta_{i})$. Similarly to the nGP specification for the elements in $\xi(t)$, we can represent the nested Gaussian Process for $\psi_{k}(t)$ with the following state equation \renewcommand{\arraystretch}{0.8} \begin{eqnarray} \left[ \begin{array}{c} \psi_{k}(t_{i+1})\\ \psi'_{k}(t_{i+1})\\ B_{k}(t_{i+1}) \end{array} \right]= \left[ \begin{array}{ccc} 1&\delta_{i}&0\\ 0 &1&\delta_{i}\\ 0 &0&1 \end{array} \right] \left[ \begin{array}{c} \psi_{k}(t_{i})\\ \psi'_{k}(t_{i})\\ B_{k}(t_{i}) \end{array} \right]+ \left[ \begin{array}{cc} 0&0\\ 1&0\\ 0&1 \end{array} \right] \left[ \begin{array}{c} \omega_{i,\psi_{k}}\\ \omega_{i,B_{k}}\\ \end{array} \right] \label{eq:14} \end{eqnarray} \renewcommand{\arraystretch}{0.5}independently for $k=1,...,K$, where $[\omega_{i,\psi_{k}}, \omega_{i,B_{k}}]^T\sim \mbox{N}_{2}(0,S_{i,k})$, with $S_{i,k}=\mbox{diag}(\sigma^2_{\psi_{k}} \delta_{i}, \sigma^2_{B_{k}} \delta_{i})$. Similarly to $\xi_{lk}(t)$ \begin{eqnarray*} [\psi_{k}(t_{1}),D^{1}\psi_{k}(t_{1}),...,D^{m-1}\psi_{k}(t_{1})]^{T} & \sim & \mbox{N}_{m}(0,\sigma^{2}_{\mu_{k}}I_{m}),\\ \ [B_{k}(t_{1}),D^{1}B_{k}(t_{1}),...,D^{n-1}B_{k}(t_{1})]^{T} & \sim & \mbox{N}_{n}(0,\sigma^{2}_{\alpha_{k}}I_{n}), \label{eq:15} \end{eqnarray*} There are two crucial aspects to highlight. Firstly, this formulation allows continuous time and an irregular grid of observations over $t$ by relating the latent states at $i+1$ to those at $i$ through the distance between $t_{i+1}$ and $t_{i}$ where $i$ represents a discrete order index and $t_{i} \in \mathcal{T}$ the time value related to the $i$th observation. Secondly, compared to \citet{Zhu:2012} our approach represents an important generalization in: (i) extending the analysis to the multivariate case (i.e. $y_{i}$ is a {\em p}-dimensional vector instead of a scalar) and (ii) accommodating locally adaptive smoothing not only on the mean but also on the time-varying covariance functions. \subsection{LAF interpretation} Model (\ref{subeq1})-(\ref{subeq2}) can be induced by marginalizing out the $K$-dimensional latent factors vector $\eta_i$, in the model \begin{eqnarray} Y_{i}=\Lambda(t_{i})\eta_{i}+\epsilon_{i}, \quad \epsilon_{i}\sim \mbox{N}_{p}(0,\Sigma_{0}) \label{eq:2} \end{eqnarray} where $\eta_{i}=\psi(t_{i})+\nu_{i}$ with $\nu_{i}\sim \mbox{N}_{K}(0,I_{K})$ and elements $\psi_{k }(t)\sim \mbox{nGP}(\sigma^2_{\psi_{k}},\sigma^2_{B_{k}})$ for $k=1,..., K$. In LAF formulation we assume moreover that the time-varying factor loadings matrix $\Lambda(t)$ is a sparse linear combination, with respect to the weights of the $p \times L$ matrix $\Theta$, of a much smaller set of continuous nested Gaussian Processes $\xi_{lk}(t)\sim \mbox{nGP}(\sigma^2_{\xi_{lk}},\sigma^2_{A_{lk}})$ comprising the $L\times K$, with $L<<p$, matrix $\xi(t)$. As a result \begin{eqnarray} \Lambda(t_{i})=\Theta\xi(t_{i}) \label{lamb} \end{eqnarray} Such a decomposition plays a crucial role in further reducing the number of nGP processes to be modeled from $p \times K$ to $L \times K$ leading to a more computationally tractable formulation in which the induced $\Gamma = \{ \mu(t),\Sigma(t), t \in \mathcal{T}\}$ follows a locally adaptive factor $\mbox{LAF}_{L,K}(\Theta, \Sigma_{0},\Sigma_{\xi} ,\Sigma_{A},\Sigma_{\psi}, \Sigma_{B})$ process where \begin{subequations} \begin{align} \mu(t_i)&=\mbox{E}(Y_{i} \ | \ t=t_i)=\Theta\xi(t_i)\psi(t_i) \label{subeq3}\\ \Sigma(t_i)&= \mbox{cov}(Y_{i} \ | \ t=t_i)=\Theta\xi(t_i)\xi(t_i)^T\Theta^T+\Sigma_{0}. \label{subeq4} \end{align} \end{subequations} There is a literature on using Bayesian factor analysis with time-varying loadings, but essentially all the literature assumes discrete-time dynamics on the loadings while our focus is instead on allowing the loadings, and hence the induced $\Gamma = \{ \mu(t),\Sigma(t), t \in \mathcal{T}\}$ processes, to evolve flexibly in continuous time. Hence, we are most closely related to the literature on Gaussian process latent factor models for spatial and temporal data; refer, for example, to \citet{Lop:2008} and \citet{Lop:2011}. In these models, the factor loadings matrix characterizes spatial dependence, with time varying factors accounting for dynamic changes. \citet{Fox:2011} instead allow the loadings matrix to vary through a continuous time stochastic process built from latent $\mbox{GP}(0,c)$ dictionary functions independently for all $l,k$, with $c$ the squared exponential correlation function having $c(x,x')=\exp(-\kappa|x-x' ||^{2}_{2})$. In our work we follow the lead of \citet{Fox:2011} in using a nonparametric latent factor model as in (\ref{eq:2})-(\ref{lamb}), but induce fundamentally different behavior on $\Gamma = \{ \mu(t),\Sigma(t), t \in \mathcal{T}\}$ by carefully modifying the stochastic processes for the dictionary functions. Note that the above decomposition of $\Gamma=\{ \mu(t),\Sigma(t), t \in \mathcal{T}\}$ is not unique. Potentially we could constrain the loadings matrix to enforce identifiability \citep{Gew:1996}, but this approach induces an undesirable order dependence among the responses (\citealp{Ag:2000}, \citealp{West:2003}, \citealp{Lop:2004}, \citealp*{Charv:2008}). Given our focus on estimation of $\Gamma$ we follow \citet{Gosh:2009} in avoiding identifiability constraints, as such constraints are not necessary to ensure identifiability of the induced mean $\mu(t)$ and covariance $\Sigma(t)$. The characterization of the class of time-varying covariance matrices $\Sigma(t)$ is proved by Lemma 2.1 of \citet{Fox:2011} which states that for $K$ and $L$ sufficiently large, any covariance regression can be decomposed as in (\ref{subeq2}). Similar results are obtained for the mean process. \subsection{Prior Specification} We adopt a hierarchical prior specification approach to induce a prior $P$ on $\Gamma = \{ \mu(t),\Sigma(t), t \in \mathcal{T}\}$ with the goal of maintaining simple computation and allowing both covariances and means to evolve flexibly over continuous time. Specifically \begin{itemize} \item{$\Gamma |\Theta, \Sigma_{0}, \Sigma_{\xi} ,\Sigma_{A},\Sigma_{\psi}, \Sigma_{B} \sim \mbox{LAF}_{L,K}(\Theta, \Sigma_{0},\Sigma_{\xi} ,\Sigma_{A},\Sigma_{\psi}, \Sigma_{B})$ } \item{Recalling the nGP assumption for the elements of $\xi(t)_{L \times K}$: $\xi_{lk}(t)\sim \mbox{nGP}(\sigma^2_{\xi_{lk}},\sigma^2_{A_{lk}})$ within LAF representation, we assume for each each element $[\Sigma_{\xi}]_{lk}$ and $[\Sigma_{A}]_{lk}$ of the $L \times K$ matrices $\Sigma_{\xi}$ and $\Sigma_{A}$ respectively, the following priors \begin{eqnarray*} \sigma^{2}_{\xi_{lk}} & \sim & \mbox{InvGa}(a_{\xi},b_{\xi}) \\ \sigma^{2}_{A_{lk}} & \sim & \mbox{InvGa}(a_{A},b_{A}) \label{eq:10} \end{eqnarray*} independently for each $(l,k)$; where $\mbox{InvGa}(a,b)$ denotes the Inverse Gamma distribution with shape $a$ and scale $b$.} \item{Similarly, the variances $[\Sigma_{\psi}]_{k}=\sigma^{2}_{\psi_{k}}$ and $[\Sigma_{B}]_{k}=\sigma^{2}_{B_{k}}$ in the state equation representation of the nGP for each $\psi_k(t) \sim \mbox{nGP}(\sigma^2_{\psi_{k}},\sigma^2_{B_{k}})$ are assumed \begin{eqnarray*} \sigma^{2}_{\psi_{k}} & \sim & \mbox{InvGa}(a_{\psi},b_{\psi}) \\ \sigma^{2}_{B_{k}} & \sim & \mbox{InvGa}(a_{B},b_{B}) \label{eq:16} \end{eqnarray*} independently for each $k$.} \item{ To address the issue related to the selection of the number of dictionary elements a shrinkage prior is proposed for $\Theta$. In particular, following \citet{Bhat:2011} we assume: \begin{eqnarray} \theta_{jl}|\phi_{jl},\tau_{l}\sim \mbox{N}(0,\phi_{jl}^{-1}\tau_{l}^{-1}) \quad \phi_{jl}\sim \mbox{Ga}(3/2,3/2) \nonumber\\ \vartheta_{1}\sim \mbox{Ga}(a_{1},1), \quad \vartheta_{h}\sim\mbox{ Ga}(a_{2},1), h \geq2, \quad \tau_{l}=\prod_{h=1}^{l}\vartheta_{h} \label{eq:12} \end{eqnarray} Note that if $a_{2}>1$ the expected value for $\vartheta_{h}$ is greater than $1$. As a result, as $l$ goes to infinity, $\tau_{l}$ tends to infinity shrinking $\theta_{jl}$ towards zero. This leads to a flexible prior for $\theta_{jl}$ with a local shrinkage parameter $\phi_{jl}$ and a global column-wise shrinkage factor $\tau_{l}$ which allows many elements of $\Theta$ being close to zero as $L$ increases.} \item{Finally for the variances of the error terms in vector $\epsilon_{i}$, we assume the usual inverse gamma prior distribution. Specifically \begin{eqnarray*} \sigma_{j}^{-2} \sim \mbox{Ga}(a_{\sigma},b_{\sigma}) \label{eq:13} \end{eqnarray*} independently for each $j=1,...,p$. } \end{itemize} \section{ Posterior Computation} For a fixed truncation level $L^{*}$ and a latent factor dimension $K^{*}$, the algorithm for posterior computation alternates between a simple and efficient simulation smoother step \citep{Durb:2002} to update the state space formulation of the nGP in LAF prior, and standard Gibbs sampling steps for updating the parametric component parameters from their full conditional distributions. \subsection{Gibbs Sampling} We outline here the main features of the algorithm for posterior computation based on observations $(y_{i},t_{i})$ for $i=1,...,T$, while the complete algorithm is provided in the Appendix. \begin{description} \item {A.} Given $\Theta$ and $\{\eta_{i}\}_{i=1}^{T}$, a multivariate version of the MCMC algorithm proposed by \citet{Zhu:2012} draws posterior samples from each dictionary element's function $\{\xi_{lk}(t_{i})\}_{i=1}^{T}$, its first order derivative $\{\xi'_{lk}(t_{i})\}_{i=1}^{T}$, the corresponding instantaneous mean $\{A_{lk}(t_{i})\}_{i=1}^{T}$, the variances in the state equations $\sigma_{\xi_{lk}}^{2}$, $\sigma_{A_{lk}}^{2}$ and the variances of the error terms in the observation equation $\sigma_{j}^{2}$ with $j=1,...,p$. \item {B.} Given $\Theta$, $\{\sigma_{j}^{-2}\}_{j=1}^{p}$, $\{ y_{i} \}_{i=1}^{T}$ and $\{ \xi(t_{i})\}_{i=1}^{T}$ we implement a block sampling of $\{\psi(t_{i})\}_{i=1}^{T}$, $\{\psi'_{k}(t_{i})\}_{i=1}^{T}$, $\{B_{k}(t_{i})\}_{i=1}^{T}$ ,$\sigma_{\psi_{k}}^{2}$, $\sigma_{B_{k}}^{2}$ and $\nu_i$ following a similar approach as in step A. \item {C.} Conditioned on $\{y_{i}\}_{i=1}^{T}$, $\{\eta_{i}\}_{i=1}^{T}$, $\{\sigma_{j}^{-2}\}_{j=1}^{p}$ and $\{\xi(t_{i})\}_{i=1}^{T}$, and recalling the shrinkage prior for the elements of $\Theta$ in (\ref{eq:12}), we update $\Theta$, each local shrinkage hyperparameter $\phi_{jl}$ and the global shrinkage hyperparameters $\tau_{l}$ following the standard conjugate analysis. \item {D. Given the posterior samples from $\Theta$, $\Sigma_{0}$, $\{ \xi(t_{i})\}_{i=1}^{T}$ and $\{\psi(t_{i})\}_{i=1}^{T}$ the realization of LAF process for $\{\mu(t_i),\Sigma(t_i), t_i \in \mathcal{T} \}$ conditioned on the data $\{y_{i}\}_{i=1}^{T}$ is \begin{eqnarray*} \mu(t_i)&=&\Theta\xi(t_i)\psi(t_i)\\ \Sigma(t_i)&= &\Theta\xi(t_i)\xi(t_i)^T\Theta^T+\Sigma_{0}. \end{eqnarray*}} \end{description} \subsection{Hyperparameter interpretation} We now focus our attention on the hyperparameters of the priors for $\sigma^{2}_{\xi_{lk}}$, $\sigma^{2}_{A_{lk}}$, $\sigma^{2}_{\psi_{k}}$ and $\sigma^{2}_{B_{k}}$. Several simulation studies have shown that the higher the variances in the latent state equations, the better our formulation accommodates locally adaptive smoothing for sudden changes in $\Gamma$. A theoretical support for this data-driven consideration can be identified in the connection between the nGP and the nested smoothing splines. It has been shown by \citet{Zhu:2012} that the posterior mean of the trajectory $U$ with reference to the problem of nonparametric mean regression under the nGP prior can be related to the minimizer of the equation \begin{eqnarray*} \frac{1}{T}\sum_{i=1}^{T}(y_{i}-U(t_{i}))^{2}+\lambda_{U}\int_{\mathcal{T}}(D^{m}U(t)-C(t))^2 dt+\lambda_{C}\int_{\mathcal{T}}(D^{n}C(t))^2 dt, \label{eq:17} \end{eqnarray*} where $C$ is the locally instantaneous function and $\lambda_{U} \in \Re^{+}$ and $\lambda_{C} \in \Re^{+}$ regulate the smoothness of the unknown functions $U$ and $C$ respectively, leading to less smoothed patterns when fixed at low values. The resulting inverse relationship between these smoothing parameters and the variances in the state equation, together with the results in the simulation studies, suggest to fix the hyperparameters in the Inverse Gamma prior for $\sigma^{2}_{\xi_{lk}}$, $\sigma^{2}_{A_{lk}}$, $\sigma^{2}_{\psi_{k}}$ and $\sigma^{2}_{B_{k}}$ so as to allow high variances in the case in which the time series analyzed are expected to have strong changes in their covariance (or mean) dynamic. A further confirmation of the previous discussion is provided by the structure of the simulation smoother required to update the dictionary functions in our Gibbs Sampling for posterior computation. More specifically, the larger the variances of $\{\omega_{i,\xi_{lk}}\}_{i=1}^{T}$, $\{\omega_{i,A_{lk}}\}_{i=1}^{T}$ and $\{\omega_{i,\psi_{k}}\}_{i=1}^{T}$, $\{\omega_{i,B_{k}}\}_{i=1}^{T}$ in the state equations, with respect to those of the vector of observations $\{y_{i}\}_{i=1}^{T}$, the higher is the weight associated to innovations in the filtering and smoothing techniques, allowing for less smoothed patterns both in the covariance and mean structures (see \citealp{Durb:2002}). In practical applications, it may be useful to obtain a first estimate of $\tilde\Gamma=\{\tilde{\mu}(t), \tilde{\Sigma}(t)\}$ to set the hyperparameters. More specifically, $\tilde{\mu}_{j}(t_{i})$ can be the output of a standard moving average on each time series $y_{j}=[y_{j1},...,y_{jT}]$, while $\tilde{\Sigma}(t_{i})$ can be obtained by a simple estimator, such as the EWMA procedure. With these choices, the recursive equation \begin{eqnarray*} \tilde{\Sigma}(t_{i})=(1-\lambda)\{[y_{i-1}-\tilde{\mu}(t_{i-1})][y_{i-1}-\tilde{\mu}(t_{i-1})]^T\}+\lambda \tilde{\Sigma}(t_{i-1}) \label{eq:18} \end{eqnarray*} become easy to implement. \subsection{Online Updating} The problem of online updating represents a key point in multivariate time series with high frequency data. Referring to our formulation, we are interested in updating an approximated posterior distribution for $\Gamma_{T+H}= \{\mu(t_{T+h}), \Sigma(t_{T+h}), h=1, ..., H\}$ once a new vector of observations $\{y_{i}\}_{i=T+1}^{T+H}$ is available, instead of rerunning posterior computation for the whole time series. Using the posterior estimates of the Gibbs sampler based on observations available up to time $T$, $\{y_{i}\}_{i=1}^{T}$, it is easy to implement (see in Appendix) a highly computationally tractable online updating algorithm which alternates between steps A, B and D outlined in the previous section for the new set of observations, and that can be initialized at $T+1$ using the one step ahead predictive distribution for the latent state vectors in the state space formulation. Note that the initialization procedure for latent state vectors in the algorithm depends on the sample moments of the posterior distribution for the latent states at $T$. As is known for Kalman smoothers (see, e.g., \citealp{Durb:2001}), this could lead to computational problems in the online updating due to the larger conditional variances of the latent states at the end of the sample (i.e., at $T$). To overcome this problem, we replace the previous assumptions for the initial values with a data-driven initialization scheme. In particular, instead of using only the new observations for the online updating, we run the algorithm for $\{y_{i}\}_{i=T-k}^{T+H}$, with $k$ small, and choosing a diffuse but proper prior for the initial states at $T-k$. As a result the distribution of the smoothed states at $T$ is not anymore affected by the problem of large conditional variances leading to better online updating performance. \section{ Simulation Studies} The aim of the following simulation studies is to compare the performance of our proposed LAF with respect to BCR, and to the models for multivariate stochastic volatility most widely used in practice, specifically: EWMA, PC-GARCH, GO-GARCH and DCC-GARCH. In order to assess whether and to what extent LAF can accommodate, in practice, even sharp changes in the time-varying means and covariances and to evaluate the costs of our flexible approach in settings where the mean and covariance functions do not require locally adaptive estimation techniques, we focus on two different sets of simulated data. The first is based on an underlying structure characterized by locally varying smoothness processes, while the second has means and covariances evolving in time through smooth processes. In the last subsection we also analyze the performance of the proposed online updating algorithm. \subsection{Simulated Data} \begin{description} \item{A. Locally varying smoothness processes:} We generate a set of $5$-dimensional observations $y_{i}$ for each $t_{i}$ in the discrete set $\mathcal{T}_{o}=\{1,2,...,100\}$, from the latent factor model in (\ref{eq:2}) with $\Lambda(t_{i})=\Theta\xi(t_{i})$. To allow sharp changes of means and covariances in the generating mechanism, we consider a $2 \times 2$ (i.e. $L=K=2$) matrix $\{\xi(t_{i})\}_{i=1}^{100}$ of time-varying functions adapted from \citet{Dono:1994} with locally-varying smoothness (more specifically we choose `bumps' functions). The latent mean dictionary elements $\{\psi(t_{i})\}_{i=1}^{100}$ are simulated from a Gaussian process $\mbox{GP}(0,c)$ with length scale $\kappa=10$, while the elements in matrix $\Theta$ can be obtained from the shrinkage prior in (\ref{eq:12}) with $a_{1}=a_{2}=10$. Finally the elements of the diagonal matrix $\Sigma_{0}^{-1}$ are sampled independently from $\mbox{Ga}(1,0.1)$. \item{B. Smooth processes: We consider the same dataset of $10$-dimensional observations $y_{i}$ with $t_{i} \in \mathcal{T}_{o}=\{1,2,...,100\}$ investigated in \citet[section~4.1]{Fox:2011}. The settings are similar to the previous with exception of $\{\xi(t_{i})\}_{i=1}^{100}$ which are $5 \times 4$ (i.e. $L=5, K=4$) matrices of smooth GP dictionary functions with length-scale $\kappa=10$.} \end{description} \subsection{Estimation Performance} \begin{description} \item{A. Locally varying smoothness processes:}\\ Posterior computation for LAF is performed by using truncation levels $L^{*}=K^{*}=2$ (at higher level settings we found that the shrinkage prior on $\Theta$ results in posterior samples of the elements in the additional columns concentrated around $0$). We place a $\mbox{Ga}(1,0.1)$ prior on the precision parameters $\sigma_{j}^{-2}$ and choose $a_{1}=a_{2}=2$. As regards the nGP prior for each dictionary element $\xi_{lk}(t)$ with $l=1,...,L^{*}$ and $k=1,...,K^{*}$, we choose diffuse but proper priors for the initial values by setting $\sigma^{2}_{\mu_{lk}}=\sigma^{2}_{\alpha_{lk}}=100$ and place an $\mbox{InvGa}(2, 10^{8})$ prior on each $\sigma^{2}_{\xi_{lk}}$ and $\sigma^{2}_{A_{lk}}$ in order to allow less smoothed behavior according to a previous graphical analysis of $\tilde{\Sigma}(t_{i})$ estimated via EWMA. Similarly we set $\sigma^{2}_{\mu_{k}}=\sigma^{2}_{\alpha_{k}}=100$ in the prior for the initial values of the latent state equations resulting from the nGP prior for $\psi_{k}(t)$, and consider $a_{\psi}=a_{B}=b_{\psi}=b_{B}=0.005$ to balance the rough behavior induced on the nonparametric mean functions by the settings of the nGP prior on $\xi_{lk}(t)$, as suggested from previous graphical analysis. Note also that for posterior computation, we first scale the predictor space to $(0,1]$, leading to $\delta_{i}=1/100,$ for $i=1,...,100$. For inference in BCR we consider the same previous hyperparameters setting for $\Theta$ and $\Sigma_{0}$ priors as well as the same truncation levels $K^{*}$ and $L^{*}$, while the length scale $\kappa$ in GP prior for $\xi_{lk}(t)$ and $\psi_{k}(t)$ has been set to 10 using the data-driven heuristic outlined in \citet{Fox:2011}. In both cases we run $50{,}000$ Gibbs iterations discarding the first $20{,}000$ as burn-in and thinning the chain every $5$ samples. As regards the other approaches, EWMA has been implemented by choosing the smoothing parameter $\lambda$ that minimizes the mean squared error (MSE) between the estimated covariances and the true values. PC-GARCH algorithm follows the steps provided by \citet{Burn:2005} with GARCH(1,1) assumed for the conditional volatilities of each single time series and the principal components. GO-GARCH and DCC-GARCH recall the formulations provided by \citet{Weide:2002} and \citet{Eng:2002} respectively, assuming a GARCH(1,1) for the conditional variances of the processes analyzed, which proves to be a correct choice in many financial applications and also in our setting. Note that, differently from LAF and BCR, the previous approaches do not model explicitly the mean process $\{ \mu(t_{i})\}_{i=1}^{100}$ but work directly on the innovations $\{ y_{i}-\mu(t_{i})\}_{i=1}^{100}$. Therefore in these cases we first model the conditional mean via smoothing spline and in a second step we estimate the models working on the innovations. The smoothing parameter for spline estimation has been set to $0.7$, which was found to be appropriate to best reproduce the true dynamic of $\{ \mu(t_{i})\}_{i=1}^{100}$. \item{B. Smooth processes:}\\ We mainly keep the same setting of the previous simulation study with few differences. Specifically, $L^{*}$ and $K^{*}$ has been fixed to $5$ and $4$ respectively (also in this case the choice of the truncation levels proves to be appropriate, reproducing the same results provided in the simulation study of \citet{Fox:2011} where $L^{*}=10$ and $K^{*}=10$). Moreover the scale parameters in the Inverse Gamma prior on each $\sigma^{2}_{\xi_{lk}}$ and $\sigma^{2}_{A_{lk}}$ has been set to $10^4$ in order to allow a smoother behavior according to a previous graphical analysis of $\tilde{\Sigma}(t_{i})$ estimated via EWMA, but without forcing the nGP prior to be the same as a GP prior. Following \citet{Fox:2011} we run $10{,}000$ Gibbs iterations which proved to be enough to reach convergence, and discarded the first $5{,}000$ as burn-in. \end{description} In the first set of simulated data, we analyzed mixing by the Gelman-Rubin procedure (see e.g. \citealp{Gel:1992}), based on potential scale reduction factors computed for each chain by splitting the sampled quantities in $6$ pieces of same length. The analysis shows slower mixing for BCR compared with LAF. Specifically, in LAF $95\%$ of the chains have a potential reduction factor lower than $1.35$, with a median equal to $1.11$, while in LAF the $95$th quantile is $1.44$ and the median equals $1.18$. Less problematic is the mixing for the second set of simulated data, with potential scale reduction factors having median equal to $1.05$ for both approaches and $95$th quantiles equal to $1.15$ and $1.31$ for LAF and BCR, respectively. \begin{figure}[t] \centering \includegraphics[height=7.3cm, width=15cm]{SMOT_LOCA.pdf} \put (-405,185) {{\footnotesize{$\Sigma_{2,2}(t_{i})$}}} \put (-265,120) {{\footnotesize{$\Sigma_{1,3}(t_{i})$}}} \put (-115,120) {{\footnotesize{$\mu_{5}(t_{i})$}}} \put (-405,80) {{\footnotesize{$\Sigma_{9,9}(t_{i})$}}} \put (-265,80) {{\footnotesize{$\Sigma_{10,3}(t_{i})$}}} \put (-40,80) {{\footnotesize{$\mu_{5}(t_{i})$}}} \caption{\footnotesize{For locally varying smoothness simulation (top) and smooth simulation (bottom), plots of truth (black) and posterior mean respectively of LAF (solid red line) and BCR (solid green line) for selected components of the variance (left), covariance (middle), mean (right). For both approaches the dotted lines represent the $95 \%$ highest posterior density intervals.}} \label{F2} \end{figure} Figure~\ref{F2} compares, in both simulated samples, true and posterior mean of the process $\Gamma = \{ \mu(t_i),\Sigma(t_i), i =1,...,100\}$ over the predictor space $\mathcal{T}_{o}$ together with the point-wise $95\%$ highest posterior density (hpd) intervals for LAF and BCR. From the upper plots we can clearly note that our approach is able to capture conditional heteroscedasticity as well as mean patterns, also in correspondence of sharp changes in the time-varying true functions. The major differences compared to the true values can be found at the beginning and at the end of the series and are likely to be related to the structure of the simulation smoother which also causes a widening of the credibility bands at the very end of the series; for references regarding this issue see \citet{Durb:2001}. However, even in the most problematic cases, the true values are within the bands of the $95\%$ hpd intervals. Much more problematic is the behavior of the posterior distributions for BCR which badly over-smooth both covariance and mean functions leading also to many $95\%$ hpd intervals not containing the true values. Bottom plots in Figure~\ref{F2} show that the performance of our approach is very close to that of BCR, when data are simulated from a model where the covariances and means evolve smoothly across time and local adaptivity is therefore not required. This happens even if the hyperparameters in LAF are set in order to maintain separation between nGP and GP prior, suggesting large support property for the proposed approach. \begin{table}[t] \centering \label{tab:1} \begin{tabular}{lcccc} &Mean &90th Quantile&95th Quantile&Max\\ \hline \multicolumn{5}{c}{Covariance $\{\Sigma(t_{i})\}$}\\ \hline EWMA&$1.37$&$2.28$& $5.49$&$85.86$\\ PC-GARCH&$1.75$&$2.49$&$6.48$&$229.50$\\ GO-GARCH&$2.40$&$3.66$&$10.32$&$173.41$\\ DCC-GARCH&$1.75$&$2.21$&$6.95$&$226.47$\\ BCR&$1.80$&$2.25$&$7.32$&$142.26$\\ LAF&$0.90$&$1.99$&$4.52$&$36.95$\\ \hline \multicolumn{5}{c}{Mean $\{\mu(t_{i})\}$}\\ \hline SPLINE&$0.064$&$0.128$& $0.186$&$2.595$\\ BCR&$0.087$&$0.185$&$0.379$&$2.845$\\ LAF&$0.062$&$0.123$&$0.224$&$2.529$\\ \hline \end{tabular} \caption{\footnotesize{LOCALLY VARYING SMOOTHNESS PROCESSES: Summaries of the standardized squared errors between true values $\{\mu(t_{i})\}_{i=1}^{100}$ and $\{\Sigma(t_{i})\}_{i=1}^{100}$ and estimated quantities $\{\hat{\Sigma}(t_{i})\}_{i=1}^{100}$ and $\{\hat{\mu}(t_{i})\}_{i=1}^{100}$ computed with different approaches.}} \label{tab:1} \end{table} \begin{table}[h!] \centering \begin{tabular}{lcccc} &Mean &90th Quantile&95th Quantile&Max\\ \hline \multicolumn{5}{c}{Covariance $\{\Sigma(t_{i})\}$}\\ \hline EWMA&$0.030$&$0.081$& $0.133$&$1.119$\\ PC-GARCH&$0.018$&$0.048$&$0.076$&$0.652$\\ GO-GARCH&$0.043$&$0.104$&$0.202$&$1.192$\\ DCC-GARCH&$0.022$&$0.057$&$0.110$&$0.466$\\ BCR&$0.009$&$0.019$&$0.039$&$0.311$\\ LAF&$0.009$&$0.022$&$0.044$&$0.474$\\ \hline \multicolumn{5}{c}{Mean $\{\mu(t_{i})\}$}\\ \hline SPLINE&$0.007$&$0.019$& $0.027$&$0.077$\\ BCR&$0.005$&$0.015$&$0.024$&$0.038$\\ LAF&$0.005$&$0.017$&$0.026$&$0.050$\\ \hline \end{tabular} \caption{\footnotesize{SMOOTH PROCESSES: Summaries of the standardized squared errors between true values $\{\mu(t_{i})\}_{i=1}^{100}$ and $\{\Sigma(t_{i})\}_{i=1}^{100}$ and estimated quantities $\{\hat{\Sigma}(t_{i})\}_{i=1}^{100}$ and $\{\hat{\mu}(t_{i})\}_{i=1}^{100}$ computed with different approaches.}} \label{tab:2} \end{table} \renewcommand{\arraystretch}{1} The comparison of the summaries of the squared errors between true process $\Gamma = \{ \mu(t_i),\Sigma(t_i), i =1,...,100\}$ and the estimated elements of $\hat\Gamma=\{\hat{\mu}(t_{i}), \hat{\Sigma}(t_{i}), i=1,\dots,100\}$ standardized with the range of the true processes $r_{\mu}=\max_{i,j}\{\mu_{k}(t_{i})\}-\min_{i,j}\{\mu_{j}(t_{i})\}$ and $r_{\Sigma}=\max_{i,j,k}\{\Sigma_{j,k}(t_{i})\}-\min_{i,j,k}\{\Sigma_{j,k}(t_{i})\}$ respectively, once again confirms the overall better performance of our approach relative to all the considered competitors. Table \ref{tab:1} shows that, when local adaptivity is required, LAF provides a superior performance having standardized residuals lower than those of the other approaches. EWMA seems to provide quite accurate estimates, but it is important to underline that we choose the optimal smoothing parameter $\lambda$ in order to minimize the MSE between estimated and true parameters, which are clearly not known in practical applications. Different values of $\lambda$ reduces significantly the performance of EWMA, which shows also lack of robustness. The closeness of the summaries of LAF and BCR in Table \ref{tab:2} confirms the flexibility of LAF even in settings where local adaptivity is not required and highlights the better performance of the two approaches with respect to the other competitors also when smooth processes are investigated. \renewcommand{\arraystretch}{.7} To better understand the improvement of our approach in allowing locally varying smoothness and to evaluate the consequences of the over-smoothing induced by BCR on the distribution of $y_{i}$ with $i=1,...,100$ consider Figure~\ref{F3} which shows, for some selected series $\{y_{ji}\}_{i=1}^{100}$ in the first simulated dataset, the time varying mean together with the point-wise $2.5\%$ and $97.5\%$ quantiles of the marginal distribution of $y_{ji}$ induced respectively by the true mean and true variance, the posterior mean of $\mu_{j}(t_{i})$ and $\Sigma_{jj}(t_{i})$ from our proposed approach and the posterior mean of the same quantities from BCR. We can clearly see that the marginal distribution of $y_{ji}$ induced by BCR is over-concentrated near the mean, leading to incorrect inferences. Note that our proposal is also able to accommodate heavy tails, a typical characteristic in financial series. \subsection{Online Updating Performance} To analyze the performance of the online updating algorithm in LAF model, we simulate $50$ new observations $\{y_{i}\}_{i=101}^{150}$ with $t_{i} \in \mathcal{T}_{o}^{*}=\{101,..., 150\}$, considering the same $\Theta$ and $\Sigma_{0}$ used in the generating mechanism for the first simulated dataset and taking the $50$ subsequent observations of the bumps functions for the dictionary elements $\{\xi(t_{i})\}_{i=101}^{150}$; finally the additional latent mean dictionary elements $\{\psi(t_{i})\}_{i=101}^{150}$ are simulated as before maintaining the continuity with the previously simulated functions $\{\psi(t_{i})\}_{i=1}^{100}$. \begin{figure}[t] \centering \includegraphics[height=10cm, width=12cm]{PRED1.pdf} \caption{\footnotesize{Plot for $4$ selected simulated series of the time-varying mean $\mu_{j}(t_{i})$ and the time-varying $2.5\%$ and $97.5\%$ quantiles of the marginal distribution of $y_{ji}$ with true mean and variance (black), mean and variance from posterior mean of LAF (red), mean and variance from posterior mean of BCR (green). Black points represent the simulated data.}} \label{F3} \end{figure} According to the algorithm described in subsection $3.3$, we fix $\Theta$, $\Sigma_{0}$, $\Sigma_{\xi}$, $\Sigma_{A}$,$\Sigma_{\psi}$ and $\Sigma_{B}$ at their posterior mean from the previous Gibbs sampler and consider the last three observations $y_{98}$, $y_{99}$ and $y_{100}$ (i.e. $k=3$) to initialize the simulation smoother in $i=101$ through the proposed data-driven initialization approach. Posterior computation shows good performance in terms of mixing, and convergence is assessed after $5{,}000$ Gibbs iterations with a small burn-in of $500$. Figure~\ref{F4} compares true mean and covariance to posterior mean of a select set of components of $\Gamma_*=\{\mu(t_{i}), \Sigma(t_{i}), i=101,...,150\}$ including also the $95\%$ hpd intervals. The results clearly show that the online updating is characterized by a good performance which allows to capture the behavior of new observations conditioning on the previous estimates. Note that the posterior distribution of the approximated mean and covariance functions tends to slightly over-estimate the patterns of the functions at sharp changes, however also in these cases the true values are within the bands of the credibility intervals. Finally note that the data-driven initialization ensures a good behavior at the beginning of the series, while the results at the end have wider uncertainty bands as expected. \begin{figure}[t] \centering \includegraphics[height=12cm, width=12cm]{UPD1.pdf} \caption{\footnotesize{Plots of truth (black) and posterior mean of the online updating procedure (solid red line) for selected components of the covariance (top), variance (middle), mean (bottom). The dotted lines represent the $95 \%$ highest posterior density intervals.}} \label{F4} \end{figure} \section{Application Study} Spurred by the recent growth of interest in the dynamic dependence structure between financial markets in different countries, and in its features during the crises that have followed in recent years, we applied our LAF to the multivariate time series of the main national stock market indices. \subsection{National Stock Indices (NSI), Introduction and Motivation} National Stock Indices represent technical tools that allow, through the synthesis of numerous data on the evolution of the various stocks, to detect underlying trends in the financial market, with reference to a specific basis of currency and time. More specifically, each Market Index can be defined as a weighted sum of the values of a set of national stocks, whose weighting factors is equal to the ratio of its market capitalization in a specific date and overall of the whole set on the same date. In this application we focus our attention on the multivariate weekly time series of the main $33$ (i.e. $p=33$) National Stock Indices from $12/07/2004$ to $25/06/2012$. Figure~\ref{F5} shows the main features in terms of stationarity, mean patterns and volatility of two selected NSI downloaded from \url{ http://finance.yahoo.com/}. The non-stationary behavior, together with the different bases of currency and time, motivate the use of logarithmic returns $y_{ji}=\log (I_{ji}/I_{ji-1})$, where $I_{ji}$ is the value of the National Stock Index $j$ at time $t_{i}$. Beside this, the marginal distribution of log returns shows heavy tails and irregular cyclical trends in the nonparametric estimation of the mean, while EWMA estimates highlight rapid changes of volatility during the financial crises observed in the recent years. All these results, together with large settings and high frequency data typical in financial fields, motivate the use of our approach to obtain a better characterization of the time-varying dependence structure among financial markets. \begin{figure}[t] \centering \includegraphics[height=12cm, width=15cm]{SUMMARY2.pdf} \put (-405,340) {USA NASDAQ} \put (-193,340) {ITALY FTSE MIB} \put (-400,250) {\footnotesize{observed time series}} \put (-187,250) {\footnotesize{observed time series}} \put (-400,135) {\footnotesize{log returns}} \put (-187,135) {\footnotesize{log returns}} \put (-400,95) {\footnotesize{conditional variances}} \put (-187,95) {\footnotesize{conditional variances}} \caption{\footnotesize{Plots of the main features of USA NASDAQ (left) and ITALY FTSE MIB (right). Specifically: observed time series (top), log-returns series (middle) with nonparametric mean estimation via $12$ week Equally Weighted Moving Average (red) in the middle, EWMA volatility estimates (bottom).}} \label{F5} \end{figure} \subsection{LAF for National Stock Index (NSI)} We consider the heteroscedastic model $y_{i}\sim \mbox{N}_{33}(\mu(t_{i}),\Sigma(t_{i}))$ for $i=1,...,415$ and $t_{i}$ in the discrete set $\mathcal{T}_{o}=\{1,2,...,415\}$, where the elements of $\Gamma=\{\mu(t_i), \Sigma(t_i), 1=1,...,415\}$, defined by (\ref{subeq3})-(\ref{subeq4}), are induced by the dynamic latent factor model outlined in \ref{eq:2} and \ref{lamb}. Posterior computation is performed by first rescaling the predictor space $\mathcal{T}_{o}$ to $(0,1]$ and using the same setting of the first simulation study, with the exception of the truncation levels fixed at $K^{*}=4$ and $L^{*}=5$ (which we found to be sufficiently large from the fact that the last few columns of the posterior samples for $\Theta$ assumed values close to $0$) and the hyperparameters of the nGP prior for each $\xi_{lk}(t)$ and $\psi_{k}(t)$ with $l=1,...,L^{*}$ and $k=1,...,K^{*}$, set to $a_{\xi}=a_{A}=a_{\psi}=a_{B}=2$ and $b_{\xi}=b_{A}=b_{\psi}=b_{B}=5 \times 10^7$ to capture also rapid changes in the mean functions according to Figure~\ref{F5}. Missing values in our dataset do not represent a limitation since the Bayesian approach allows us to update our posterior considering solely the observed data. We run $10{,}000$ Gibbs iterations with a burn-in of $2{,}500$. Examination of trace plots of the posterior samples for $\Gamma=\{\mu(t_{i}), \Sigma(t_{i}), i=1,...,415\}$ showed no evidence against convergence. Posterior distributions for the variances in Figure~\ref{F6} demonstrate that we are clearly able to capture the rapid changes in the dynamics of volatility that occur during the world financial crisis of $2008$, in early $2010$ with the Greek debt crisis and in the summer of $2011$ with the financial speculation in government bonds of European countries together with the rejection of the U.S. budget and the downgrading of the United States rating. \begin{figure}[t] \centering \includegraphics[height=9cm, width=15cm]{POST2.pdf} \put (-405,254) {USA NASDAQ} \put (-188,254) {ITALY FTSE MIB} \put (-401,145) {\footnotesize{log returns}} \put (-185,145) {\footnotesize{log returns}} \put (-401,110) {\footnotesize{conditional variances}} \put (-185,110) {\footnotesize{conditional variances}} \caption{\footnotesize{Top: Plot for $2$ NSI, respectively USA NASDAQ (left) and ITALY FTSE MIB (right), of the log returns (black) and the time-varying estimated mean $\{\hat{\mu}_{j}(t_{i})\}_{i=1}^{415}$ together with the time-varying $2.5\%$ and $97.5\%$ quantiles (red) of the marginal distribution of $y_{ji}$ from LAF. Bottom: posterior mean (black) and $95\%$ hpd (dotted red) for the variances $\{\Sigma_{jj}(t_{i})\}_{i=1}^{415}$.}} \label{F6} \end{figure} Moreover, the resulting marginal distribution of the log returns induced by the posterior mean of $\mu_{j}(t)$ and $\Sigma_{jj}(t)$, shows that we are also able to accommodate heavy tails as well as mean patterns cycling irregularly between slow and more rapid changes. Important information about the ability of our model to capture the evolution of world geo-economic structure during different finance scenarios are provided in Figures~\ref{F7} and~\ref{F8}. \begin{figure}[t] \centering \includegraphics[height=10cm, width=13cm]{R_NEW.pdf} \put (-320,260) {LAF} \put (-320,119) {BCR} \caption{\footnotesize{Black line: For USA NASDAQ median of correlations with the other $32$ NSI based on posterior mean of $\{\Sigma(t_{i})\}_{i=1}^{415}$. Red lines: $25\%$, $75\%$ (dotted lines) and $50\%$ (solid line) quantiles of correlations between USA NASDAQ and European countries (without considering Greece and Russia which present a specific pattern). Green lines: $25\%$, $75\%$ (dotted lines) and $50\%$ (solid line) quantiles of correlations between USA NASDAQ and the countries of Southeast Asia (Asian Tigers and India). Timeline: (A) burst of U.S. housing bubble; (B) risk of failure of the first U.S. credit agencies (Bear Stearns, Fannie Mae and Freddie Mac); (C) world financial crisis after the Lehman Brothers' bankruptcy; (D) Greek debt crisis; (E) financial reform launched by Barack Obama and EU efforts to save Greece (the two peaks represent respectively Irish debt crisis and Portugal debt crisis); (F) worsening of European sovereign-debt crisis and the rejection of the U.S. budget; (G) crisis of credit institutions in Spain and the growing financial instability of the Eurozone.}} \label{F7} \end{figure} From the correlations between NASDAQ and the other National Stock Indices (based on the posterior mean $\{\hat{\Sigma}(t_{i})\}_{i=1}^{415}$ of the covariances function) in Figure~\ref{F7}, we can immediately notice the presence of a clear geo-economic structure in world financial markets (more evident in LAF than in BCR), where the dependence between the U.S. and European countries is systematically higher than that of South East Asian Nations (Economic Tigers), showing also different reactions to crises. Plots at the top of the Figure~\ref{F8} confirms the above considerations showing how Western countries exhibit more connection with countries closer in terms of geographical, political and economic structure; the same holds for Eastern countries where we observe a reversal of the colored curves. As expected, Russia is placed in a middle path between the two blocks. A further element that our model captures about the structure of the markets is shown in the plots at the bottom of Figure~\ref{F8}. The time-varying regression coefficients obtained from the standard formulas of the conditional normal distribution based on the posterior mean of $\Gamma=\{\mu(t_{i}), \Sigma(t_{i}) , i=1,...,415\}$ highlight clearly the increasing dependence of European countries with higher crisis in sovereign debt and Germany, which plays a central role in Eurozone as expected. \begin{figure}[t] \centering \includegraphics[height=10cm, width=15cm]{REG.pdf} \put (-410,265) {GERMANY DAX30} \put (-268,265) {CHINA SSE Composite} \put (-125,265) {RUSSIA RTSI Index} \put (-410,123) {ITALY} \put (-268,123) {SPAIN} \put (-125,123) {GREECE} \caption{\footnotesize{Top: For $3$ selected stock market indices, plot of the median of the correlation based on posterior mean of $\{\Sigma(t_{i})\}_{i=1}^{415}$ with the other $32$ world stock indices (black), the European countries without considering Greece and Russia (red) and the Asian Tigers including India (green). Bottom: For $3$ of the European countries more subject to sovereign debt crisis, plot of $25th$, $50th$ and $75th$ quantiles of the time-varying regression parameters based on posterior mean $\{\hat{\Sigma}(t_{i})\}_{i=1}^{415}$ with the other countries (black) and Germany (red).}} \label{F8} \end{figure} The flexibility of the proposed approach and the possibility of accommodating varying smoothness in the trajectories over time, allow us to obtain a good characterization of the dynamic dependence structure according with the major theories on financial crisis. Top plot in Figure~\ref{F7} shows how the change of regime in correlations occurs exactly in correspondence to the burst of the U.S. housing bubble (A), in the second half of $2006$. Moreover we can immediately notice that the correlations among financial markets increase significantly during the crises, showing a clear international financial contagion effect in agreement with other theories on financial crisis (see, e.g., \citealp{Bai:1999}, and \citealp{Cl:2009}). As expected the persistence of high levels of correlation is evident during the global financial crisis between late-2008 and end-2009 (C), at the beginning of which our approach also captures a sharp variation in the correlations between the U.S. and Economic Tigers, which lead to levels close to those of Europe. Further rapid changes are identified in correspondence of Greek crisis (D), the worsening of European sovereign-debt crisis and the rejection of the U.S. budget (F) and the recent crisis of credit institutions in Spain together with the growing financial instability Eurozone (G). Finally, even in the period of U.S. financial reform launched by Barack Obama and EU efforts to save Greece (E), we can notice two peaks representing respectively Irish debt crisis and Portugal debt crisis. Note also that BCR, as expected, tends to over-smooth the dynamic dependence structure during the financial crisis, proving to be not able to model the sharp change in the correlations between USA NASDAQ and Economic Tigers during late-2008, and the two peaks representing respectively Irish and Portugal debt crisis at the beginning of 2011. \subsection{National Stock Indices, Updating and Predicting} The possibility to quickly update the estimates and the predictions as soon as new data arrive, represents a crucial aspect to obtain quantitative informations about the future scenarios of the crisis in financial markets. To answer this goal, we apply the online updating algorithm presented in subsection $3.3$, to the new set of weekly observations $\{y_{i}\}_{i=416}^{422}$ from $02/07/2012$ to $13/08/2012$ conditioning on posterior estimates of the Gibbs sampler based on observations $\{y_{i}\}_{i=1}^{415}$ available up to $25/06/2012$. We initialized the simulation smoother algorithm with the last $8$ observations of the previous sample. Plots at the top of Figure~\ref{F9} show, for $3$ selected National Stock Indices, the new observed log returns $\{y_{ji}\}_{i=416}^{422}$ (black) together with the mean and the $2.5\%$ and $97.5\%$ quantiles of the marginal distribution (red) and conditional distribution (green) of $y_{ji}|y_{i}^{-j}$ with $y_{i}^{-j}=\{y_{qi},q \neq j\}$. We use standard formulas of the multivariate normal distribution based on the posterior mean of the updated $\Gamma_*=\{\mu(t_{i}), \Sigma(t_{i}), i=416,...,422\}$ after $5{,}000$ Gibbs iterations with a burn in of $500$. This is sufficient for convergence based on examining trace plots of the time-varying mean and covariance matrices. From these results, we can clearly notice the good performance of our proposed online updating algorithm in obtaining a characterization for the distribution of new observations. Also note that the multivariate approach together with a flexible model for the mean and covariance, allow for significant improvements when the conditional distribution of an index given the others are analyzed. \begin{figure}[t] \centering \includegraphics[height=10cm, width=15.5cm]{pred2.pdf} \put (-422,275) {USA NASDAQ} \put (-275,275) {INDIA BSE30} \put (-128,275) {FRANCE CAC40} \put (-422,130) {\footnotesize{prediction: method (a)}} \put (-275,130) {\footnotesize{prediction: method (b)}} \put (-128,130) {\footnotesize{prediction: method (c)}} \caption{\footnotesize{Top: For $3$ selected NSI, respectively USA NASDAQ (left), INDIA BSE30 (middle) and FRANCE CAC40 (right), plot of the observed log returns (black) together with the mean and the $2.5\%$ and $97.5\%$ quantiles of the marginal distribution (red) and conditional distribution given the other $32$ NSI (green) based on the posterior mean of $\Gamma_*=\{\mu(t_{i}), \Sigma(t_{i}), i=416,...,422\}$ from the online updating procedure for the new observations from $02/07/2012$ to $13/08/2012$. Bottom: boxplots of the one step ahead prediction errors for the $33$ NSI, where the predicted values are respectively: (a) unconditional mean $\{\tilde{y}_{i+1}\}_{i=415}^{421}=0$, (b) marginal mean of the one step ahead predictive distribution using the online updating procedure for $\{\tilde{y}_{i+1|i}\}_{i=415}^{421}$, (c) conditional mean given the log returns of the other $32$ NSI at $i+1$ of the one step ahead predictive distribution using the online updating procedure for $\{\tilde{y}_{i+1|i}\}_{i=415}^{421}$. Predictions for (b) and (c) are induced by the posterior mean of $\{\mu(t_{i+1|i}), \Sigma(t_{i+1|i}), i=415,..,421 \}$ of our LAF.}} \label{F9} \end{figure} To obtain further informations about the predictive performance of our LAF, we can easily use our online updating algorithm to obtain $h$ step-ahead predictions for $\Gamma_{T+H|T}=\{\mu(t_{T+h|T}), \Sigma(t_{T+h|T}), h=1,...,H\}$. In particular, referring to \citet{Durb:2001}, we can generate posterior samples of $\Gamma_{T+H|T}$ merely by treating $\{y_{i}\}_{i=T+1}^{T+H}$ as missing values in the proposed online updating algorithm. Here, we consider the one step ahead prediction (i.e. $H=1$) problem for the new observations. More specifically, for each $i$ from $415$ to $421$, we update the mean and covariance functions conditioning on informations up to $t_{i}$ through the online algorithm and then obtain the predicted posterior distribution for $\Sigma(t_{i+1|i})$ and $\mu(t_{i+1|i})$ by adding to the sample considered for the online updating a last column $y_{i+1}$ of missing values. Plots at the bottom of Figure~\ref{F9}, show the boxplots of the one step ahead prediction errors for the $33$ NSI obtained as the difference between the predicted value $\tilde{y}_{j,i+1|i}$ and, once available, the observed log return $y_{j,i+1}$ with $i+1=416,...,422$ corresponding to weeks from $02/07/2012$ to $13/08/2012$. In (a) we forecast the future log returns with the unconditional mean $\{\tilde{y}_{i+1}\}_{i=415}^{421}=0$, which is what is often done in practice under the general assumption of zero mean, stationary log returns. In (b) we consider $\tilde{y}_{i+1|i}=\hat{\mu}(t_{i+1|i})$, the posterior mean of the one step ahead predictive distribution of $\mu(t_{i+1|i})$, obtained from the previous proposed approach after $5{,}000$ Gibbs iteration with a burn in of $500$. Finally in (c) we suppose that the log returns of all National Stock Indices except that of country $j$ (i.e., $y_{j,i+1}$) become available at $t_{i+1}$ and, considering $y_{i+1|i}\sim N_{p}(\hat{\mu}(t_{i+1|i}),\hat{\Sigma}(t_{i+1|i}))$ with $\hat{\mu}(t_{i+1|i})$ and $\hat{\Sigma}(t_{i+1|i})$ posterior mean of the one step ahead predictive distribution respectively for $\mu(t_{i+1|i})$ and $\Sigma(t_{i+1|i})$, we forecast $\tilde{y}_{j,i+1}$ with the conditional mean of $y_{j,i+1}$ given the other log returns at time $t_{i+1}$. Comparing boxplots in (a) with those in (b) we can see that our model allows to obtain improvements also in terms of prediction. Furthermore, by analyzing the boxplots in (c) we can notice how our ability to obtain a good characterization of the time-varying covariance structure can play a crucial role also in improving forecasting, since it enters into the standard formula for calculating the conditional mean in the normal distribution. \section{ Discussion} In this paper, we have presented a continuous time multivariate stochastic process for time series to obtain a better characterization for mean and covariance temporal dynamics. Maintaining simple conjugate posterior updates and tractable computations in moderately large $p$ settings, our model increases significantly the flexibility of previous approaches as it captures sharp changes both in mean and covariance dynamics while accommodating heavy tails. Beside these key advantages, the state space formulation enables development of a fast online updating algorithm particularly useful for high frequency data. The simulation studies highlight the flexibility and the overall better performance of LAF with respect to the models for multivariate stochastic volatility most widely used in practice, both when adaptive estimation techniques are required, and also when the underlying mean and covariance structures do not show sharp changes in their dynamic. The application to the problem of capturing temporal and geo-economic structure between the main financial markets demonstrates the utility of our approach and the improvements that can be obtained in the analysis of multivariate financial time series with reference to (i) heavy tails, (ii) locally adaptive mean regression, (iii) sharp changes in covariance functions, (iii) high dimensional dataset, (iv) online updating with high frequency data (v) missing values and (vi) predictions. Potentially further improvements are possible using a stochastic differential equation model that explicitly incorporates prior information on dynamics.
\section{Introduction} \label{Introduction} Type II radio bursts, especially in the decameter--hectometric (D--H) and kilometer (km) wavelength range, are thought to be caused by the electron beam accelerated by CME--driven shocks \citep[e.g.][]{SheeleyJr:1985taa,1998GeoRL..25.2493R,1999GeoRL..26.1573B}. Assuming that the type II radio burst was excited at the shock front, the speed of the shock could be obtained from the observed frequency drift rate of the type II radio burst based on a coronal--density model \citep[e.g.][]{2001A&A...377..321V,Gopalswamy:2002jq,2002A&A...396..673V,2003AIPC..679..152R,2004A&A...413..753V}. This method is widely used to study and forecast the propagation of shocks \citep[e.g.][]{Dryer:1984ua,1990SoPh..129..387S,Fry:2003vh,2007ApJ...663.1369R}. \citet{Shen:2007ww} established a method to derive shock strength from the observations based on the assumption that type II radio bursts were generated from the nose of shocks. However, whether or not type II radio bursts originate from the nose of shocks is still an open question. As a high--density region, the streamer was thought to be a place where strong shocks easily form \citep{2008ApJ...687.1355E}. Thus, the shock--streamer interaction region was discussed as a possible source region of type II radio bursts \citep[e.g.][]{Cho:2005jv,Cho:2007da,Cho:2008cf,Cho:2011be}. Using the coronal--density distribution obtained from the Mauna Loa Solar Observatory ( MLSO ) MK4 polarization map, \citet{Cho:2007da} studied the relationship between a metric type II burst and a CME. They found that the metric type II burst was generated at the interface of the CME flank and the streamer. Further, \citet{Cho:2008cf} checked the source regions of 19 metric type II radio bursts and found that both of the front and the CME--streamer interaction regions are the possible source regions for the metric type II radio bursts. In addition, the shock and streamer interaction could also affect the spectrum of the type II radio burst. Recently, \citet{2012ApJ...750..158K} and \citet{2012ApJ...753...21F} reported two clear cases where the spectrum of type II radio bursts varied during the interaction between the shocks and the streamers. However, the discussion of the source regions of the type II radio bursts in the decameter--hectometric (D--H) wavelength range is still lacking. The spectrum of RAD2 on \textit{Wind/Waves} \citep{Bougeret:1995ce} is from 1 to 14 MHz corresponding to the corona region from $\approx$2 $R_\odot$ to $\approx$9 $R_\odot$. The C2 camera of the \textit{Large Angle and Spectrometric COronagraph} (LASCO; \citet{Brueckner:1995cb}) onboard the \textit{SOlar and Heliospheric Observatory} (SOHO; \citet{Domingo:1995}) provides the imaging observations from 1.5 $R_\odot$ to 6 $R_\odot$. Thus, the combination of \textit{Wind/Waves} and SOHO/LASCO--C2 observations could be used to study the source regions and the variation of type II radio bursts at the D--H wavelength range. In this article, we check source regions of two well--observed D--H type II radio bursts. The method applied in this study is introduced in Section 2. In Sections 3 and 4, two typical D--H type II events and their source regions are studied. Conclusions and discussions are then given in the last section. \section{Method}\label{method} In this work, the source region of the D--H type II radio burst is obtained from the combined analysis of the \textit{Wind/Waves} and SOHO/LASCO observations. The detailed method is described as follows: \begin{enumerate}[i)] \item Previous results suggest that type II radio bursts especially in the D--H and longer wavelength range are caused by the CME--driven shocks \citep{1998GeoRL..25.2493R,1999GeoRL..26.1573B}. Recently, \citet{Vourlidas:2003kn} and \citet{2009ApJ...693..267O} found that shocks could be directly observed in coronagraph images. We use SOHO/LASCO--C2 observations to identify the position of the shock front, called S$_{\mbox{shock}}$ hereafter. It is thought to be the possible source region of type II bursts. \item The fundamental frequency of a type II burst is related to the background electron density by \citep{1982GAM....21.....P}: \begin{eqnarray} N_e=(\frac{f_{pe}(Hz)}{8.98\times10^3})^2\mbox{cm}^{-3}\label{eq1}. \end{eqnarray} Therefore, the electron density of the source regions of the type II could be obtained from the radio burst dynamic spectrum. Using the pb\_inverter.pro procedure in Solar Software (SSW\footnote{\href{http://www.lmsal.com/solarsoft/}{http://www.lmsal.com/solarsoft/}}), the polarized--brightness observations from SOHO/LASCO could be used to get the background electron density distribution. The pb\_inverter.pro procedure uses the pB inversion derivation obtained by \citet{1950BAN....11..135V}. A polynomial fit of the form r$^{-n}$ is applied to the pB image for a single position angle to get the electron density distribution (see the introduction of the `pb\_inverter.pro' in the SSW). Thus, the possible regions, called as S$_\rho$, which can generate the type II radio bursts at the observed frequency range at the time of shock observed, can be determined. In addition, considering a 2\% uncertainty in the brightness observations (Vourlidas, 2012, private communicate), a 2\% uncertainty in the obtained electron density was applied to find the S$_\rho$. \item The overlap region of the shock front (S$_{\mbox{shock}}$) and the derived density region (S$_\rho$) is defined as the source region of the type II radio burst. \end{enumerate} Based on the method described above, to determine the source region of a D--H type II radio burst, we need the polarized--brightness image and the direct imaging observations of the shock from SOHO/LASCO and the D--H type II radio--burst observation from \textit{Wind/Waves}. Thus, we select events based on the following criteria: \begin{enumerate}[i)] \item A clear type II radio burst was recorded by \textit{Wind/Waves}. \item The burst was caused by a limb CME with clear shock signatures in the LASCO--C2 field of view. We require limb CMEs because the polarized--brightness image represents the background coronal--density distribution near the plane of the sky, and the shock driven by a limb CME should have the most clear signatures in coronagraph. \end{enumerate} Conforming to these two criteria, two well--observed events were found for study in this article. \section{7 March 2011 Event} Figure \ref{mar7_lasco} shows the SOHO/LASCO observations before and after the onset of this CME. On 7 March 2011, a limb CME originating from N24W59 was first observed by SOHO/LASCO--C2 at 20:00 UT. The orange $\ast$ symbols in panel (b) of Figure \ref{mar7_lasco} show the possible front positions of this CME at 20:00 UT. Using the GCS model \citep{Thernisien:2006ke,2009SoPh..256..111T,Thernisien:2011jy}, \citet{chenglong:2012wa} obtained the speed of this CME as 2115 $\pm$ 136 km\ s$^{-1}$ in the three--dimensional space. This is a very fast CME with a speed much faster than the local Alfv\'en speed, and therefore LASCO--C2 only captured three images of the CME. We can expect that this CME drove a shock when it propagated in the corona. Seen from panel (c) and (d) in Figure \ref{mar7_lasco}, obvious shock signatures ahead of the main body of the CME could be identified. The orange $\ast$ symbols in panel (c) and (d) of Figure \ref{mar7_lasco} mark the shock front at two instants of time. As we described in Section 2, the shock front, S$_{\mbox{shock}}$, is thought to be the possible source region of the associated type II radio burst. Figure \ref{mar7_wave} shows the \textit{Wind/Waves} observations from 7 March 2011 19:50 UT to 21:00 UT. From Figure \ref{mar7_wave}, an obvious type II radio burst could be identified. The vertical dotted--dashed white lines in Figure \ref{mar7_wave} indicate the times of the shock recorded by SOHO/LASCO--C2. Seen from this figure, the signature of the type II radio burst at 20:12 UT is very weak. Near 20:22 UT, this type II radio burst became stronger, and lasted for about ten minutes. The white asterisks show the minimum and maximum fundamental frequency of this D--H type II radio burst at the time of 20:24 UT, which are 4.6 and 7.4 MHz corresponding to the electron density of $2.6\times10^5$ cm$^{-3}$ and $6.8\times10^5$ cm$^{-3}$, respectively, based on Equation \ref{eq1}. \begin{figure} \center \noindent\includegraphics[width=\hsize]{event1.lasco.eps} \caption{The SOHO/LASCO--C2 observations for the 7 March 2011 event. Panel (a) shows the polarized--brightness image. Panels (b) -- (d) show the CME at different times.} \label{mar7_lasco} \end{figure} \begin{figure} \center \noindent\includegraphics[width=\hsize]{event1.waves.eps} \caption{The \textit{Wind/Waves} observations of 7 March 2011 19:50 UT to 21:00 UT.} \label{mar7_wave} \end{figure} Figure \ref{mar7_rho} shows the background electron--density distribution obtained from the polarized image at 08:58 UT. The white regions in Figure \ref{mar7_rho} are caused by the unsuccessful determination of the density. It is clearly seen in the figure that the electron density varies significantly with position angle. This suggests that a simple one--dimensional density model may not reflect the real condition. The regions, i.e. S$_\rho$, in which the electron density falls in the range from $2.6\times10^5$ to $6.8\times 10^5$ cm$^{-3}$, are indicated in red. The type II radio burst near 20:24 UT were probably generated from these region. The position of S$_{\mbox{shock}}$ based on coronagraph observations is overplotted with yellow $\ast$ on Figure 3. As we discussed in Section \ref{method}, the source region of a type II radio burst is the overlap region between the S$_{\mbox{shock}}$ and S$_\rho$. For this event at 7 March 2011 20:24 UT, the source region is located in the region indicated by the blue rectangle. The white + symbols in Figure \ref{mar7_rho} show the boundary of the streamer, which is determined from the SOHO/LASCO image before the onset of the CME as indicated by the green + in Figure \ref{mar7_lasco}(a). It is found that the source of this type II radio burst is located at the shock--streamer interaction region. This result suggests that the type II radio burst at the D--H frequency range might also originate from the shock--streamer interaction region, similar to the metric type II radio bursts \citep{Cho:2007da,Cho:2008cf}. In addition, at 20:12 UT, the shock was also very clear in the SOHO/LASCO--C2 image (Figure 1c), but, the type II signature was much weaker than that near 20:24 UT. We suggest that such a difference is probably attributed to the degree of interaction between the shock and the streamer. From Figure \ref{mar7_lasco}(c), it seems that the shock did not fully interact with the streamer at 20:12 UT. But, at 20:24 UT as shown in Figure \ref{mar7_lasco}(d), a part of the CME--driven shock was propagating in the streamer obviously. The interaction between the streamer and the shock enhanced the shock and then the enhanced shock increased the intensity of the type II radio burst. \begin{figure} \center \noindent\includegraphics[width=\hsize]{event1.rho.eps} \caption{The electron--density distribution obtained from the polarized--brightness image from SOHO/LASCO for the 7 March 2011 event. }\label{mar7_rho} \end{figure} \section{9 May 2011 event} This CME burst from the east limb of the solar disk. SOHO/LASCO--C2 observed it since 21:24 UT. It was a fast limb CME with a projected speed of 1318 km\ s$^{-1}$. Figure \ref{may9_lasco} shows the SOHO/LASCO observations before and after the onset of this CME. It is found that the shock structure ahead of the CME could be well observed and identified at 21:24 UT, 21:36 UT and 21:48 UT based on SOHO/LASCO--C2 observations. The orange $\ast$ symbols in panels (b) -- (d) show the shock front at three instants of time, which are defined as S$_{\mbox{shock}}$. A D--H type II radio burst associated with this CME is shown in Figure \ref{may9_wave}. It started at $\approx$21:15 UT. At 21:24 UT, the type II radio burst was very weak and is difficult to identify. The half of the frequency of its harmonic component, which varied from 2.5 MHz to 3.3 MHz, is used as the fundamental frequency. After 21:24 UT, the strength of this D--H type II increased. This radio burst reached its strongest phase near 21:48 UT. At 21:36 UT, the fundamental frequency varied from 1.9 MHz to 2.7 MHz. Figure \ref{may9_rho} shows the electron--density distribution, which is derived from the polarized--brightness image recorded at 14:58 UT. The regions of S$_\rho$ at 21:24 UT and 21:36 UT are indicated by the red color in panels (a) and (b), respectively. Similar to Figure \ref{mar7_rho}, the yellow $\ast$ symbols indicate S$_{\mbox{shock}}$. The source regions of this type II event at the two instants of time were located in the regions enclosed by the blue rectangles in Figure \ref{may9_rho}. \begin{figure} \center \noindent\includegraphics[width=\hsize]{event2.lasco.eps} \caption{The SOHO/LASCO--C2 observations for the 9 May 2011 event as for Figure \ref{mar7_lasco}. }\label{may9_lasco} \end{figure} The white + symbols in Figure \ref{may9_rho} show the boundary of the streamer as same as the green + in Figure \ref{may9_lasco}(a). It is found that the source regions of this event at different time also located in the shock--streamer interaction regions. It confirms the conclusion that the shock--streamer interaction region might be the source region of a D--H type II radio burst. Seen from the Figure \ref{may9_lasco} (b) and (c), it is found that the interaction between the shock and the streamer may start near 21:24 UT. After that, the shock further interacted with the streamer. During this phase, the observed D--H type II radio burst enhanced continuously as shown in Figure \ref{may9_wave}. Thus, the increase of the intensity of this radio burst was probably caused by the enhancement of the shock during its interaction with the streamer. \begin{figure} \center \noindent\includegraphics[width=\hsize]{event2.waves.eps} \caption{The \textit{Wind/Waves} observations from 9 May 2011 21:00 UT to 22:30 UT, similar to Figure \ref{mar7_wave}.}\label{may9_wave} \end{figure} \begin{figure} \center \noindent\includegraphics[width=\hsize]{event2.rho.eps} \caption{ The electron--density distribution obtained from the polarized--brightness image from SOHO/LASCO for the 9 May 2011 event. }\label{may9_rho} \end{figure} \section{Conclusion} In this work, the source regions of two well--observed D--H type II radio bursts are checked based on SOHO/LASCO--C2 and \textit{Wind/Waves} observations. It is found that the source regions of these two D--H type II radio bursts probably located in the shock--streamer interaction regions, which is the same as the source regions of metric type II radio bursts \citep{Cho:2007da,Cho:2008cf,Cho:2011be}. In addition, by analyzing the intensity variation of these two D--H type II radio bursts, we suggest that the shocks were enhanced during their interaction with the streamer. Such enhancement of shocks would increase the intensity of the radio bursts. These results indicate that the shock--streamer interaction region could also be one of the main source region of the D--H type II radio burst. It should be noted that the background density in a streamer (or the flank of a shock) is quite different from that near the nose of a shock. Thus, to calculate the shock speed based on the frequency drift rate of type II radio burst, a detailed analysis of where a radio burst is generated from should be done first. As we described in Section \ref{method}, the background density obtained from the SOHO/LASCO polarized--brightness image is an important factor in our method. Unfortunately, there is only one polarized image taken each day in each telescope for most period of the SOHO mission\footnote{\href{http://lasco-www.nrl.navy.mil/index.php?p=content/retrieve/products}{http://lasco-www.nrl.navy.mil/index.php?p=content/retrieve/products}}. Solar eruptions, especially the large CMEs, would significantly influence the background density. Thus, we choose only the events in which no large CME events occurred between the time of the polarized--brightness image recorded and the type II radio burst. In addition, clear type II radio--burst observations and clear shock signatures in SOHO/LASCO observations are needed in this method. Combined with these selection criteria, the number of events could be studied is limited. In a future work, the method developed by \citet{Hayes:2001ii} could be used to obtain the background electron density based on the total--brightness images, and more events could be studied. It should be noted that only projection observations were used in this work. Recently, some methods were developed to obtain the CME's parameters \citep[e.g.][]{Thernisien:2006ke, 2009SoPh..256..111T; Thernisien:2011jy; 2012arXiv1203.3261F}, background electron density in the corona \citep[e.g.][]{Frazin:2010fc} and the streamer structure \citep[e.g.][]{2010ApJ...710....1M} in three--dimensional space. Thus, the three--dimensional source region of the radio burst could be further checked by applying various three--dimensional models. We acknowledge the use the SOHO/LASCO and \textit{Wind/Waves} observations. The SOHO/LASCO data used here are produced by a consortium of the Naval Research Laboratory (USA), Max-Planck-Institut f\"{u}r Aeronomie (Germany), Laboratoire d'Astrophysique de Marseille (France), and the University of Birmingham (UK). SOHO is a mission of international cooperation between ESA and NASA. This work is supported by the CAS Key Research Program (KZZD-EW-01), grants from the 973 key project 2011CB811403, NSFC 41131065, 40904046, 41274173, 40874075, and 41121003, CAS the 100-talent program, KZCX2-YW-QN511 and startup fund, and MOEC 20113402110001 and the fundamental research funds for the central universities (WK2080000031 and WK2080000007).
\section{Introduction} We consider a wireless sensor network deployed for the purpose of monitoring a common phenomenon. The sensors transmit their observations to a fusion center in a cooperative manner, so as to conserve the overall (energy/power) resources available in the network.\xdef\@thefnmark{}\@footnotetext{This research was partially supported by the National Science Foundation under Grant No. 0925854 and the Air Force Office of Scientific Research under Grant No. FA-9550-10-C-0179.} In the widely researched area of \emph{distributed estimation} \cite{Rib06},\cite{Cui07}, the sensors are coordinated to ensure that, without communicating with one-another, they collectively maximize the quality of inference at the FC. In the amplify-and-forward approach to distributed estimation, the sensors linearly scale (based on the energy allocated) their observations while communicating with the FC. Since no coding across time is required and no non-linear processing is required at the nodes, amplify-and-forward techniques are operationally simple and enjoy widespread usage in the literature. Early application of the amplify-and-forward technique in distributed estimation was explored in \cite{Cui07} and \cite{Xiao08}, where orthogonal and coherent multiple access channels (MAC) were considered. In an orthogonal MAC setting, each sensor has its own channel for communication with the FC while in the coherent MAC scenario, all the sensors coherently form a beam into a common channel which is then received by the FC. Other approaches to distributed estimation consider rate constraint in transmission, where the sensor nodes are required to quantize their observations before transmission to the FC, examples include \cite{Rib06} and more recently \cite{KarTSP12}. Some recent studies \cite{Fang09},\cite{KarISIT12} have demonstrated significant improvement over the distributed framework by allowing the sensor nodes to share their observations with other neighboring nodes prior to transmission to the FC. This act of sharing observations is referred to as \emph{spatial collaboration}. In an orthogonal MAC setting with a fully connected network, it has been shown in \cite{Fang09} that it is optimal to compute the estimates in the network and use the best available channel to transmit the estimated parameter. In a recent work that considered a coherent MAC channel for communication with the FC \cite{KarISIT12}, we presented an extension of the amplify-and-forward framework that allowed spatial collaboration in a partially connected network topology. It was observed that even a sparsely connected network was able to realize a performance which was very close to that of a fully connected network. This is due to the fact that in an amplify-and-forward framework, the observation noise is also amplified along with the signal, thereby significantly increasing the energy required for transmission. Spatial collaboration, in effect, smooths out the observation noise, thereby improving the quality of the signal that is transmitted to the FC using the same energy resources. In this paper, we explore the potential of collaborative estimation further by making two significant contributions. First, though it was observed earlier that even a moderately connected network performs almost as well as a fully connected network, no analytical results were presented. In this paper, we extend our previous work by analyzing the estimation performance for partially connected collaboration networks. Though the analysis of arbitrary network topologies is a difficult problem, we derive the estimation performance for a family of structured network topologies, namely the $Q$-cliques. We demonstrate that the insights obtained from the structured topology apply approximately to two practical topologies, namely the nearest neighbor and random geometric graphs, of similar connectivity. Given a particular network topology, we investigate two different energy allocation schemes for data transmission. In addition to the optimal energy-allocation (EA) scheme as derived in \cite{KarISIT12}, we also consider the suboptimal but easy-to-implement equal energy-allocation scheme, where neighboring observations are simply averaged to mitigate the observation noise. For both of these schemes, we derive the performance in a closed form in the asymptotic domain where the number of nodes is large and the overall transmission capacity of the network is held constant. These results offer insights into the relationship between estimation performance and problem parameters like channel and observation gains, prior uncertainty and extent of spatial collaboration. The collaborative estimation problem has so far been analyzed in the single-snapshot context, where energy-constrained spatial sampling is performed at one particular instant and the inference is made using those samples. In our second contribution in this paper, we extend the problem formulation to consider power-constrained inference of a random process, where the goal is to estimate the process for \emph{all} time instants. In contrast to the simple snapshot framework, this involves obtaining multiple samples in time and computing the filtered estimates for any desired time instants, including time instants where observation samples are not available. Since collection of each sample involves the expenditure of energy resources, the appropriate constraint in this situation is the energy spent per unit time (or power). A key parameter here is the sampling frequency, the choice of which affects the overall estimation performance. Note that a higher sampling frequency usually means that one can better capture the temporal variations. However, with a power constraint, less energy is available for the collection of each of those samples, which would result in more noisy samples. This trade-off is investigated in the context of a Gaussian random process with exponential covariance, where it turns out that a higher sampling frequency \emph{always} results in better estimates. \begin{figure}[htb] \centering \includegraphics[width=0.99} \newcommand{\figszb}{0.8 \columnwidth]{cooperative_schematic.pdf} \caption{Wireless sensor network performing collaborative estimation.} \label{fig:schematic} \end{figure} \section{Problem Formulation} We first consider the single snapshot estimation problem, where the dimension of time is ignored. We will extend the discussion to time varying Gaussian process later in Section \ref{sec:prob:form:OU}. The single snapshot framework is depicted in Figure \ref{fig:schematic}. The parameter to be estimated, $\theta$ (written without any time subscript to signify the single snapshot nature of the problem), is assumed to be a zero-mean Gaussian source with prior variance $\eta^2$. Different noisy versions of $\theta$ are observed by $N$ sensors. The observation vector is $\bo x=[x_1,\ldots,x_N]$ where $x_n=h_n \theta+\epsilon_n$, with $h_n$ and $\epsilon_n$ denoting the observation gain and measurement noise respectively. The measurement noise variables $\{\epsilon_n\}_{n=1}^N$ are assumed to be independent and identically distributed (iid) Gaussian random variables with zero mean and variance $\sigma^2$. Let the availability of collaboration links be represented by the adjacency matrix $\bo A$, where $A_{nm}=1$ (or $A_{nm}=0$) implies that node $n$ has (or does not have) access to the observation of node $m$. Define an $\bo A$-sparse matrix as one for which non-zero elements may appear only at locations $(n,m)$ for which $A_{nm}=1$. The set of all $\bo A$-sparse matrices is denoted by $\mathcal S_A$. Corresponding to an adjacency matrix $\bo A$ and an $\bo A$-sparse matrix $\bo W$, we define \emph{collaboration} in the network as individual nodes being able to linearly combine local observations from other collaborating nodes \begin{align} z_n=\sum_{m\in \mathcal A(n)} \bo W_{nm}x_m, \label{def:zn} \end{align} where $\mathcal A(n)\triangleq\left\{m:\; \bo A_{nm}=1 \right\}$, without any further loss of information. In effect, the network is able to compute a one-shot spatial transformation of the form $\bo z=\bo W \bo x$. In practice, this transformation is realizable when any two neighboring sensors are close enough to ensure reliable information exchange. Note that, when $\bo W$ is restricted to be diagonal (in other words, when $\bo A=\bo I$), the problem reduces to the amplify-and-forward framework for distributed estimation, which is widely used in the literature \cite{Cui07},\cite{Xiao08},\cite{Fang09} due to its simplicity in implementation and provably optimal information theoretic properties for simple networks \cite{Gastpar03}. The transformed observations $\{z_n\}_{n=1}^N$ are transmitted to the FC through a coherent MAC channel, so that the received signal is $y=\bo g^T \bo z +u$, where $\bo g$ and $u$ describe the channel gains and the channel noise respectively. The channel noise $u$ is assumed to be Gaussian distributed with zero mean and variance $\xi^2$. The FC receives the noise-corrupted signal $y$ and computes an estimate of $\theta$. Since $y$ is a linear Gaussian random variable conditioned on $\theta$, \begin{align} \begin{split} \theta &\sim \mathcal N(0,\eta^2),\mbox{ and} \\ y|\theta &\sim \mathcal N \left(\underbrace{\bo g^T \bo W \bo h}_{\begin{smallmatrix} \triangleq \mu \; \text{(net gain)} \es} \theta,\, \; \underbrace{\bo g^T \bo W \bo \Sigma \bo W^T \bo g +\xi^2}_{\begin{smallmatrix} \triangleq \zeta^2 \; \text{(net noise variance)} \es}\right), \end{split} \label{linear:model} \end{align} the minimum-mean-square-error (MMSE) estimator $\widehat \theta=\mathbb E\left[ \theta|y \right]$ is the optimal fusion rule. From estimation theory (for details the reader is referred to \cite{Kay93}), the MMSE estimator and resulting distortion $D_{\bo W}$ is given by \begin{align} \widehat \theta &=\frac{1}{1+\frac{\zeta^2}{\eta^2 \mu^2}} \frac{y}{\mu},\mbox{ and } \frac{1}{D_{\bo W}}=\frac{1}{\eta^2}+J_{\bo W}, \; J_{\bo W}=\frac{\mu^2}{\zeta^2}, \label{JW} \end{align} where the quantity $J_{\bo W}$ is the Fisher Information and $\mu$ and $\zeta^2$ are the net gain and net noise variance as defined in Equation \eqref{linear:model}. The cumulative transmission energy required to transmit the transformed observations $\bo z$ is \begin{equation} \begin{aligned} \mathcal E_{\bo W}&=\mathbb E[\bo z^T \bo z]=\text{Tr }\left[\bo W \bo E_{\textsf x} \bo W^T\right], \mbox{ where} \\ \bo E_{\textsf x} &\triangleq \mathbb E[\bo x \bo x^T] = \eta^2 \bo h\bo h^T+\bo \Sigma. \label{def:PW} \end{aligned} \end{equation} \subsection{Collaboration strategies} Note that the quantities $\mu$, $\zeta^2$ and, therefore, the distortion $D$ (equivalently $J$) and also the transmission energy $\mathcal E$ depend on the choice of the collaboration matrix $\bo W$. As indicated earlier, we explore two strategies to determine $\bo W$, namely 1) optimal and 2) equal energy-allocation (EA) strategies, that stem from two different engineering considerations. In the optimal EA strategy, we assume that the FC knows the channel and observation gains and also the collaboration topology precisely. In such a situation, the FC can compute the optimal collaboration matrix subject to a cumulative transmission energy constraint \begin{align} \mbox{(Optimal EA)}\quad \bo W_\textsf{opt}= \arg \min_{\bo W \in \mathcal S_A} D_{\bo W}, \; \mbox{ s.t. } \mathcal E_{\bo W} \le \mathcal E, \label{EA:opt} \end{align} and communicate the corresponding weights $\bo W_\textsf{opt}$ to the sensor nodes via a separate and reliable control channel. The exact form of $\bo W_\textsf{opt}$ and corresponding $J_\textsf{opt}$ were derived in \cite{KarISIT12} and are briefly described as follows. \begin{result:opt}[Optimal single-snapshot estimation, \cite{KarISIT12}] \label{result:opt:lbl} Let $L$ be the cardinality of $\bo A$, which is also the number of non-zero collaboration weights. In an equivalent representation, construct $\bo w \in \mathbb R^L$ by concatenating those elements of $\bo W$ that are allowed to be non-zero. Accordingly, define the $L\times L$ matrix $\bo \Omega$ and $L\times N$ matrix $\bo G$ such that the identities \begin{align} \text{Tr }\left[\bo W \bo E_{\textsf x} \bo W^T\right]=\bo w^T \bo \Omega \bo w, \mbox{ and }\bo g^T \bo W =\bo w^T \bo G, \end{align} are satisfied. Then the optimal Fisher Information is, \begin{align} \begin{split} &J_{\textsf{opt}}=\bo h^T\left(\bo \Sigma+\bo \Gamma/\mathcal E_\xi \right)^{-1} \bo h, \mbox{ where} \\ &\; \mathcal E_\xi \triangleq \mathcal E/\xi^2, \mbox{ and } \bo \Gamma \triangleq \left(\bo G^T \bo \Omega^{-1}\bo G\right)^{-1}, \end{split} \label{J:opt:actual} \end{align} which is achieved when the collaboration weights are $\bo w_{\textsf{opt}}=\kappa \bo \Omega^{-1}\bo G \bo \Gamma\left(\bo \Sigma+ \bo \Gamma/\mathcal E_\xi \right)^{-1} \bo h$, with the scalar $\kappa$ chosen to satisfy $\bo w_{\textsf{opt}}^T\bo \Omega \bo w_{\textsf{opt}}=\mathcal E$. $\bo W_\textsf{opt}$ is the matrix equivalent of $\bo w_\textsf{opt}$. \end{result:opt} When either the FC is computationally limited or reliable control channels are not available, the optimal EA strategy cannot be implemented. In these situations, one reasonable way of assigning transmission energy and collaboration weights at each node may be the equal EA strategy, where all the sensors are allocated equal transmission energy (namely $\frac{\mathcal E}{N}$). In addition, the $n$th sensor equally weighs all the observations from its neighbors where the weights (say $\{d_n\}_{n=1}^N$) are chosen to satisfy $\mathbb E\left[ z_n^2 \right]=\frac{\mathcal E}{N}$. Note from \eqref{def:zn} that $z_n= d_n\sum_{m\in \mathcal A(n)} (h_m\theta+ \epsilon_m)$. Consequently, \begin{equation} \begin{aligned} &\mbox{(Equal EA)} \quad \left[\bo W_{\textsf{eq}}\right]_{nm}=\left\{ \begin{array}{rl} d_n, & \mbox{if } m\in \mathcal A(n) \\ 0, & \mbox{else} \end{array} \right.,\\ &\quad d_n=\sqrt{\frac{\mathcal E/N}{ \left(\sum_{m\in \mathcal A(n)} h_m \right)^2\eta^2 +| \mathcal A(n) | \sigma^2}}, \end{aligned} \label{EA:eq} \end{equation} where $| \mathcal A(n) |$ denotes the number of neighbors of $n$. The Fisher Information corresponding to the equal EA strategy is simply $J_\textsf{eq}\triangleq J_{\bo W_\textsf{eq}}$, which can be obtained by applying Equation \eqref{JW}. Once the collaboration strategy (namely, either optimal or equal EA) is chosen and a cumulative operating energy $\mathcal E$ is specified, the resulting distortion performance (FI-s $J_\textsf{opt}$ or $J_\textsf{eq}$) depends on the following problem parameters, 1) signal prior, measurement noise and channel noise, which were assumed to be Gaussian distributed with zero mean and variances $\eta^2$, $\sigma^2 \bo I_N$ and $\xi^2$ respectively, 2) observation and channel gains, and 3) the collaboration topology. To obtain analytical expressions for FI-s, it is clear that we need to make further simplifying assumptions on the observation/channel gains (which will be discussed in Section \ref{sec:snapshot}) and also the topology for collaboration. \begin{figure}[htb] \centering \includegraphics[width=\figszb \columnwidth]{collab_Qconn.pdf} \caption{Example of $Q$-cliques.} \label{fig:Qconn} \end{figure} \subsection{Partially connected networks} \label{sec:partial} With the goal to investigate partially connected collaboration topologies, we adopt the following methodology. Intuitively, we expect the distortion to decrease as the network becomes more connected (since it adds more degrees of freedom) and it is our aim to obtain asymptotic limits that explicitly reflect the effect of \emph{connectedness}. Since the analysis of arbitrary topologies is difficult, we derive our analytical results for a structured collaboration topology that consists of several fully-connected clusters (or cliques) of finite size $Q$, as illustrated in Figure \ref{fig:Qconn}. To be precise, if $N=KQ$, we have $\bo A=\bo I_K \otimes \left( \bo 1_Q \bo 1_Q^T \right)$. Since all of the nodes in a $Q$-clique network are $(Q-1)$-connected, the performance of this special topology may serve as an approximation to other topologies where the average number of neighbors per node is $(Q-1)$. To demonstrate the efficacy of this approximation, we will compare the analytical results for $Q$-clique networks with numerical results for two practical collaboration topologies, namely 1) nearest-neighbor (NN) topology and 2) random geometric graphs (RGG) \cite{Freris10}. For the NN-topology, a sensor receives the observations from its nearest $Q-1$ neighbors. For the RGG topology, a sensor collaborates with all other sensors that are located within a circle of radius $r$ with the sensor at the center. The expected number of neighbors, which is a function of $r$, can be derived using geometric arguments, thereby enabling comparison with an equivalent $Q$-clique topology. Examples of these two topologies are illustrated in Figures \ref{fig:topology:rgg} and \ref{fig:topology:nn} for a network with $N=20$ nodes. It may be noted that unlike the NN topology, all collaboration links of an RGG topology are bidirectional by definition. \begin{figure}[htb] \centering \subfigure[Random geometric graph with radius $r=0.2$ (total 44 links)]{ \includegraphics[width=\figszb \columnwidth]{collab_ideal_rgg.pdf} \label{fig:topology:rgg} } \subfigure[Nearest neighbor topology with $Q=3$ (total 40 links) ]{ \includegraphics[width=\figszb \columnwidth]{collab_ideal_nn2.pdf} \label{fig:topology:nn} } \caption{Example of two practical collaboration topologies for a $N=20$-node network. Bidirectional links are shown without arrows.} \label{fig:topology} \end{figure} \subsection{Ornstein-Uhlenbeck process } \label{sec:prob:form:OU} We now bring in the dimension of time, and consider the problem of power-constrained estimation of time-varying signals. In order to model the temporal dynamics, we assume that the signal of interest is a stationary zero-mean Gaussian random process $\theta_t$ with exponential covariance function \begin{align} \mathbb E\left[ \theta_{t_1}, \theta_{t_2} \right]=\eta^2 e^{-\left( |t_1-t_2| \right)/\tau}, \label{cov:OU} \end{align} where $\eta^2$ and $\tau$ represent the magnitude and temporal variation of the parameter respectively. Note that $\tau\rightarrow 0$ implies that the signal changes very rapidly, while $\tau\rightarrow \infty$ means that the signal is constant over time. Such a process (with covariance parameterized by $\eta^2$ and $\tau$) is widely used in the literature \cite{Niu10} due to its ability to model a time varying Gaussian process while providing a relatively simple framework for analysis. This process is also sometimes known as the Ornstein-Uhlenbeck (OU) process. \begin{figure}[htb] \centering \includegraphics[width=0.99} \newcommand{\figszb}{0.8 \columnwidth]{collab_sampling_def.pdf} \caption{Periodically sampled Ornstein-Uhlenbeck process.} \label{fig:sampling} \end{figure} Let $P$ denote the power constraint in the network. We assume that the OU process is sampled periodically with the period $T$ (see Figure \ref{fig:sampling}), which implies that a total of $\mathcal E=PT$ energy units is available for each sampling instant. Let the spatial sampling at each instant be performed in a manner similar to the single-snapshot framework discussed earlier, namely through collaboration and coherent amplify-and-forward beamforming. Let the signal received by the FC at instant $t$ be denoted as $y_t$, so that the entire observed sequence can be represented by the following infinite-dimensional vector $\bo y\triangleq [\ldots,y_{-T}, y_0, y_T,y_{2T},\ldots ]'$. From the linear Gaussian model $y|\theta$ in \eqref{linear:model} and subsequent description of Fisher Information $J$ in \eqref{JW}, the spatial sampling process can be abstracted (via an appropriate scaling) through the additive model $y_t=\theta_t+v_t$, where the \emph{aggregate} noise $v_t\sim \mathcal N\left(0,\frac{1}{J}\right)$ summarizes the uncertainty due to $\left\{ \epsilon_{n,t} \right\}_{n=1}^N$ (measurement noise at sensors) and $u_t$ (channel noise). We assume $\epsilon_{n,t}$ and $u_t$ to be temporally white, from which it follows that $v_t$ is temporally white as well. In vector notations, \begin{align} \bo y=\bo \theta +\bo v, \quad \bo v\sim \mathcal N \left(0,\frac{1}{J} \bo I\right), \label{y:theta:v} \end{align} where $\bo \theta \triangleq [\ldots,\theta_{-T}, \theta_0, \theta_T,\theta_{2T},\ldots]'$ and $\bo v \triangleq [\ldots,v_{-T}, v_0, v_T, v_{2T},\ldots ]'$. Having observed $\bo y$, the MMSE estimator of $\theta_t$ (the value of OU process at any instant $t$) is given by the conditional expectation (refer to \cite{Kay93} for details) \begin{align} \widehat \theta_t=\mathbb E[\theta_t|\bo y] = \bo R_{\theta_t \bo y'} \bo R_{\bo y \bo y'}^{-1} \bo y, \label{E:theta:t} \end{align} where $\bo R_{\theta_t \bo y'}\triangleq \mathbb E\left[ \theta_t \bo y'\right]$ and $R_{\bo y \bo y'}\triangleq \mathbb E\left[ \bo y \bo y'\right]$. Moreover, the variance of $\widehat \theta_t$ is given by \begin{align} \text{Var}\left(\theta_t|\bo y\right) = \eta^2-\bo R_{\theta_t \bo y'} \bo R_{\bo y \bo y'}^{-1} \bo R_{\theta_t \bo y}. \label{var:theta:t} \end{align} Since $y_t$ is sampled periodically at instants $t=kT,\, k\in \mathbb Z$ and $\bo y$ contains infinite elements in both time directions, the conditional variance $\text{Var}\left(\theta_t|\bo y\right)$ is also expected to be periodic in time, i.e., $\text{Var}\left(\theta_t|\bo y\right)=\text{Var}\left(\theta_{t+kT}|\bo y\right),\; \forall t,k$. Hence any interval of length $T$, say $t\in[0,T]$ is sufficient for analyzing the conditional variance \eqref{var:theta:t}. Since we are interested in estimating the OU process at \emph{all} time instants, the quality of inference has to be summarized by a metric that is independent of time $t$. We consider two such performance metrics, first of which is the average variance \begin{align} \text{Avar}(T) \triangleq \frac{1}{T}\int_0^T \text{Var}\left(\theta_t|\bo y\right) \,\mathrm{d} t. \label{def:avar} \end{align} Note that average variance depends on the sampling period $T$, which is made explicit by the argument. However, there may be situations when the sampling period $T$ is also subject to design. In this case, we need a metric that is independent of $T$ as well. In this situation, we may use the performance metric to be the limiting value \begin{align} \text{Var}_0 \triangleq \min_T \max_{t\in[0,T]}\text{Var}\left(\theta_t|\bo y\right), \label{def:var0} \end{align} which assumes that we select the sampling period $T$ in a manner that minimizes the worst-case conditional variance for all time. We would consider both the metrics \eqref{def:avar} and \eqref{def:var0} in this paper. \section{Main Results} \subsection{Single snapshot estimation} \label{sec:snapshot} As motivated earlier, we consider an $N$-sensor network, the collaboration topology of which consists entirely of $Q$-cliques, where $Q$ is a finite integer (see Figure \ref{fig:Qconn}). Let $N=K Q$, which ensures that there is an integral number of such cliques. Let the total energy available in the network be $\mathcal E$, which is finite. We consider the asymptotic limit when the network is large ($N\rightarrow \infty$) but the transmission capacity of the equivalent Multiple-Input-Single-Output (MISO) channel is kept finite. We assume that the random variables $\{\widetilde g_n\}_{n=1}^N$ (which can be thought of as \emph{unnormalized} channel gains) are iid realizations from the pdf $f_{\widetilde g}(\cdot)$ and the channel gains are $ g_n=\frac{1}{\sqrt{N}} \widetilde g_n$ so as to ensure that the transmission capacity remains the same even as the number of nodes increase\footnote{Note that the channel capacity of the equivalent MISO channel is $\frac{1}{2}\log\left(1+\frac{\mathcal E \| \bo g \|^2}{\xi^2} \right)$ and that $\lim_{N\rightarrow \infty} \| \bo g \|^2=\mathbb E\left[ \widetilde g^2 \right]$ from the law of large numbers.}, thereby enabling a fair comparison of networks of various sizes. Without such a scaling, the transmission capacity would increase to infinity (and the resulting distortion would be driven down to zero) as the number of nodes increase, which is a trivial regime to consider. Let $J_\textsf{opt}$ and $J_{\textsf{eq}}$ denote the asymptotic limits of the Fisher Information $J_{\bo W}$ corresponding to the optimal \eqref{EA:opt} and equal energy-allocation strategies \eqref{EA:eq} respectively. The following results provide closed form expressions for these limits. \begin{result:J}[Fisher Information for $Q$-clique topology] \label{result:J:lbl} \begin{subequations} \begin{align} \mbox{(Optimal EA) }\quad J_\textsf{opt}&=\frac{ \mathcal E }{\eta^2} \frac{\mathbb E\left[ \widetilde g^2 \right]}{\xi^2} \left(1-H_Q \right), \mbox{ and}\label{J:opt} \\ \mbox{(Equal EA) } \quad J_{\textsf{eq}}&=\frac{ \mathcal E }{\eta^2} \frac{ \left( \mathbb E\left[ \widetilde g \right]\right)^2}{\xi^2 } \frac{1}{1+R_Q} \label{J:eq}, \end{align} \end{subequations} where $H_Q$ and $R_Q$ are defined as \begin{subequations} \begin{align} H_Q&=\mathbb E\left[\frac{1}{1+\frac{\eta^2}{\sigma^2} \left(h_1^2+\cdots+h_Q^2 \right)} \right], \mbox{ and} \label{HQ}\\ R_Q&=\frac{1}{Q \left( \mathbb E\left[ h \right] \right)^2} \left(\text{Var}\left[ h \right]+\frac{\sigma^2}{\eta^2} \right) \label{RQ}, \end{align} \end{subequations} respectively. \end{result:J} The proof of Theorem \ref{result:J:lbl} is skipped here due to lack of space and can be found in an extended version of this paper \cite{Kar12CollaborativeArxiv}. A few remarks due to Theorem \ref{result:J:lbl} are in order. \emph{Special Cases: } In general, the equal EA scheme is suboptimal, i.e., $J_\textsf{opt} \ge J_{\textsf{eq}}$. However, the two energy allocation schemes are asymptotically equivalent when $\text{Var}[h]$ and $\text{Var}[\widetilde g]$ are both zero, which is the case when the network is homogeneous, i.e., $\bo h= h_0 \bo 1$ and $\tbo g= \widetilde g_0 \bo 1$ (say). For such a network, \begin{align} J_\textsf{opt}=J_{\textsf{eq}}=\frac{\mathcal E \widetilde g_0^2}{\xi^2 \eta^2} \frac{1}{1+\frac{1}{Q} \frac{\sigma^2}{\eta^2 h_0^2} }. \end{align} \emph{Explicit expressions for Rayleigh distributed gains: } The evaluation of \eqref{J:opt} is, in general, hindered by the computation of $H_Q$, which involves the computation of a $Q$-dimensional integral. However, if the observation gains are Rayleigh\footnote{A Rayleigh distributed random variable $x$ with parameter $\alpha$ has a probability density function $\textsf{Rayleigh}(x;\alpha)=\frac{x}{\alpha^2} \exp\left( -\frac{x^2}{2 \alpha^2} \right)$ for $x\in[0,\infty)$. The first two moments are $\mathbb E\left[ x \right]=\alpha\sqrt{\frac{\pi}{2}}$ and $\mathbb E\left[x^2\right]=2\alpha^2$ respectively.} distributed $f_h(h)=\textsf{Rayleigh}(h;\alpha_h)$, we can show (the derivation is relegated to \cite{Kar12CollaborativeArxiv}) that \begin{align} H_Q=\frac{(-1)^{Q-1}\lambda^Q \exp (\lambda) \mathcal Ei(\lambda) -\sum_{i=0}^{Q-2} i! (-\lambda)^{Q-1-i}}{(Q-1)!}, \label{HQ:ray} \end{align} where $\lambda \triangleq \frac{\sigma^2}{2 \alpha_h^2 \eta^2}$ and $\mathcal Ei(z)\triangleq \int_{z}^\infty \exp(-t)/t \,\mathrm{d} t$ is the exponential integral function. It immediately follows that $H_1=\lambda^Q \exp (\lambda) \mathcal Ei(\lambda)$, which corresponds to the distributed case ($Q=1$), and $H_Q \approx \frac{\lambda}{Q-1}$ for large values of $Q$. In our numerical simulations, we would consider Rayleigh distributed channel and observation gains, for which \eqref{HQ:ray} will be useful. \subsection*{Simulation results} Theorem \ref{result:J:lbl} is important since it provides a framework to evaluate the estimation performance for partially connected collaboration topologies. Though \eqref{J:opt} and \eqref{J:eq} are accurate indicators of performance for a structured network consisting only of $Q$-cliques, it is of interest to see how this insight applies for more complicated topologies. Towards that goal, we simulate the nearest-neighbor and random geometric graph topologies as described in Section \ref{sec:partial}. We consider a network with $N=10^4$ nodes, which is large enough to demonstrate convergent behavior. We consider $\eta^2=1$, $\xi^2=1$ $f_h(h)=\textsf{Rayleigh}(h;1)$, $f_{\widetilde g}(\widetilde g)=\textsf{Rayleigh}(\widetilde g;1)$ and two values of observation noise variance, namely $\sigma^2=1$ and $\sigma^2=2$. The operating energy is fixed at $\mathcal E=0.7$. This choice of $\mathcal E$ is made to reflect an operating region where substantial performance gain is possible through spatial collaboration. \begin{figure}[htb] \centering \subfigure[Fixed number of nearest neighbors ]{ \includegraphics[width=0.99} \newcommand{\figszb}{0.8 \columnwidth]{collab_asymp_nn.pdf} \label{fig:asymp:nn} } \subfigure[Random geometric graph]{ \includegraphics[width=0.99} \newcommand{\figszb}{0.8 \columnwidth]{collab_asymp_rgg.pdf} \label{fig:asymp:rgg} } \caption{Energy-constrained estimation with single snapshot spatial sampling.} \label{fig:asymp} \end{figure} Numerical results obtained through Monte-Carlo simulations of both the optimal and equal EA strategies with varying degrees of spatial collaboration are illustrated in Figures \ref{fig:asymp:nn} and \ref{fig:asymp:rgg} for the NN and RGG topologies respectively. For the $(Q-1)$-nearest-neighbor case, the theoretical results corresponding to an equivalent problem with $Q$-cliques are juxtaposed. It is observed from Figure \ref{fig:asymp:nn} that the performance of the two schemes are almost identical. For a random geometric graph, we consider that all the $N$ sensors are randomly spread in a unit square. If the radius of collaboration is $r$, it follows that the expected number of neighbors is approximately $\widetilde Q = N \pi r^2$, and that this approximation is more accurate for large values of $r$. Hence in Figure \ref{fig:asymp:rgg}, the theoretical results corresponding to an equivalent problem with $\widetilde Q$-cliques are juxtaposed. It is observed that the theoretical approximations compare favorably with the Monte-Carlo simulations, although they are less accurate compared to the $(Q-1)$-nearest-neighbor case. From the two examples given above, it is observed that only a small number of collaboration links are needed to achieve near-connected performance. In particular, the distortion performance is seen to saturate as early as $Q\gtrapprox 20$, though a fully connected network would imply $Q=N=10^4$ connections per node. This demonstrates the efficacy of spatial collaboration as an approach to enhance estimation performance beyond distributed networks. \subsection{Time varying process estimation} \label{sec:OU} In this subsection, we compute the conditional variance of the OU process given the vector of periodically sampled observations. Towards computing \eqref{var:theta:t}, we begin by describing the matrix $\bo R_{\bo y \bo y'}$ and vector $\bo R_{\theta_t \bo y'}$. The covariance matrix of the sampled parameter values $\bo \theta$ takes the shape of the well known stationary matrix (e.g., \cite{Niu10}, \cite{Kar12}), \begin{align} \mathbb{E}[\bo \theta \bo \theta'] &= \eta^2\bo C, \; \bo C \triangleq \begin{bmatrix} 1 & \rho & \rho^2 & \hdots & \cdot \\ \rho & 1 & \ddots & \ddots & \vdots \\ \rho^2 & \ddots & \ddots & \ddots & \rho^2 \\ \vdots & \ddots & \ddots & 1 & \rho \\ \cdot & \hdots & \rho^2 & \rho & 1 \\ \end{bmatrix}, \label{R:theta} \end{align} where $\rho \triangleq e^{-T/\tau}$. The structured matrix $\bo C$ in Equation \eqref{R:theta}, is often referred to as the Kac--Murdock--Szeg$\ddot{o}$ matrix in the literature. From the additive model \eqref{y:theta:v}, it follows that \begin{align} \bo R_{\bo y \bo y'}=\left( \bo I+ \eta^2 J \bo C \right)/J. \end{align} Similarly, the following expression for $\bo R_{\theta_t \bo y'}$ follows directly from the definition $\varrho = e^{-t/\tau}$, where $t\in[0,T]$, \begin{align} \mathbb{E}[\theta_t \bo y'] &= \eta^2 \left[ \ldots, \rho^2\varrho,\rho\varrho,\varrho, \rho/\varrho,\rho^2/\varrho,\ldots\right]. \end{align} With the help of the above descriptions of $\bo R_{\bo y \bo y'}$ and $\bo R_{\theta_t \bo y'}$, computing \eqref{var:theta:t} involves inverting the matrix $\bo I+ \eta^2 J \bo C$ (the asymptotic closed form expression for such an inverse was introduced in \cite{Kar12}) followed by a quadratic product. The resulting algebra is involved but straightforward. We relegate details of the derivation to \cite{Kar12CollaborativeArxiv} and state the result below. \begin{result:var:theta}[Variance of OU process estimates] \label{result:var:theta:lbl} \begin{align}\begin{split} \text{Var}\left(\theta_t|\bo y\right)&=\frac{\eta^2\left[1+\eta^2 J\rho' \left\{ 1- \left(\frac{\varrho-\rho/\varrho}{1-\rho} \right)^2 \right\} \right]}{ \sqrt{ \left( \eta^2 J+\rho' \right) \left( \eta^2 J+1/\rho' \right) } }, \\ \mbox{where } \rho' &\triangleq \frac{1-\rho}{1+\rho}, \rho=e^{-T/\tau}, \varrho=e^{-t/\tau}, t\in[0,T]. \label{res:var:theta} \end{split}\end{align} \end{result:var:theta} Equation \eqref{res:var:theta} provides the closed form estimation variance of an OU process at any instant $t\in[0,T]$, provided that power constrained noisy samples are observed periodically with period $T$. In addition to $\rho$, the quantity $J$ also depends on the sampling period $T$ through the energy-FI (Fisher Information) relation $J=\frac{c PT}{\eta^2}$, which follows from Theorem \ref{result:J:lbl} by using $\mathcal E=P T$ and defining \begin{equation} c \triangleq \begin{cases} \frac{\mathbb E\left[ \widetilde g^2 \right] \left(1-H_Q \right)}{\xi^2} & \mbox{ for Optimal EA},\\ \frac{\mathbb (E\left[ \widetilde g \right])^2}{\xi^2 \left(1+R_Q \right)} & \mbox{ for Equal EA}. \end{cases} \label{def:c} \end{equation} In the following discussions, we illustrate the power-constrained estimation of an OU process and show how the instantaneous variance \eqref{res:var:theta} can be used to compute other performance measures described in Section \ref{sec:prob:form:OU}, namely a) average variance, $\text{Avar}(T)$ and b) min-max performance limit, $\text{Var}_0$. \subsection*{Simulation results} \begin{figure}[htb] \centering \subfigure[Effect of sampling period on aggregate noise and subsequent filtering]{ \includegraphics[width=0.99} \newcommand{\figszb}{0.8 \columnwidth]{collab_OUproc_eg.pdf} \label{fig:OUproc:eg} } \subfigure[Instantaneous variance for various sampling periods]{ \includegraphics[width=0.99} \newcommand{\figszb}{0.8 \columnwidth]{collab_OUproc_var.pdf} \label{fig:OUproc:var} } \caption{Power-constrained estimation of OU process } \vspace{-0.2in} \label{fig:OUproc} \end{figure} In Figure \ref{fig:OUproc:eg}, we visualize the estimation of an OU process with stationary variance $\eta^2=1$ and covariance drop-off parameter $\tau=1$s. We simulate a total duration of $T_\textsf{obs}=30$s, during which we consider sampling the same process using two different sampling periods, $T=0.75$s (top) and $T=3$s (bottom). The sampling noise sequence $\{ v_{kT} \}$, which is an abstraction of the spatial data aggregation, is simulated as independent zero-mean Gaussian random variables with variance $\frac{1}{2.5 T}$ for the two different sampling periods. The inverse relation $\text{Var}\left(v_{kT}\right) \propto \frac{1}{T}$ is due to the fact that $v_{kT}$ represents a noise with variance $\frac{1}{J}$ and the Fisher Information $J=\frac{c P T}{\eta^2}$ as per Theorem \ref{result:J:lbl}. The constant $2.5$ (which represents the quantity $\frac{cP}{\eta^2}$) was chosen so as to produce a visible contrast between the sampling errors corresponding to the chosen sampling periods. As can be seen in Figure \ref{fig:OUproc:eg}, the samples are obtained almost without any noise for $T=3$s (bottom). The circles representing noisy samples align almost on top of the thin line that represents the path of the OU process. The samples are, however, significantly noisy for $T=0.75$s (top), since less energy is available per sampling duration. This is evidenced by the circles lying significantly distant from the OU process path. The filtered estimates $\widehat \theta_t=\mathbb E\left[ \theta_t|\{ y_{kT} \} \right]$ are obtained by applying \eqref{E:theta:t} and are shown by the bold lines. When compared visually, the filtered estimates appear more accurate in the case of smaller sampling period (top). This observation is justified by plotting the steady state variance, as obtained from Theorem \ref{result:var:theta:lbl}, in Figure \ref{fig:OUproc:var} for various sampling periods $T=\{0.1,0.75,1.5,3\}$. Though the best-case variance (occurring at $t=kT$) is higher for smaller sampling periods, the worst-case variance (occurring at $t=(k+0.5)T$) goes down as the OU process is sampled more frequently. Because the power is kept constant, the variance converges to a finite value (rather than vanishing) for small values of $T$. Since the gap between the worst-case and best-case scenarios reduces with $T$, the limiting variance ($\approx 0.4$, annotated as $\text{Var}_0$) is flat with respect to time. From Equation \eqref{res:var:theta}, the limiting variance can be derived precisely to be \begin{align} \text{Var}_0\triangleq \lim_{T\rightarrow 0} \text{Var}\left(\theta_t|\bo y\right) = \frac{\eta^2}{\sqrt{1+2 P\tau c}}, \label{var0} \end{align} which also means that $\text{Var}_0$ trivially satisfies \begin{align} \min_T \max_{t\in[0,T]}\text{Var}\left(\theta_t|\bo y\right)=\text{Var}_0, \end{align} thereby answering the question of min-max performance limit as posed earlier in Equation \eqref{def:var0}. The following result is obtained by substituting in \eqref{var0} the value of constant $c$ (see \eqref{def:c}), thereby stating explicitly how the performance limit depends on channel conditions and collaboration topology. \begin{result:var0}[Min-max performance limits] \label{result:var0:lbl} \begin{equation} \text{Var}_0 = \begin{cases} \eta^2 \bigg/ \sqrt{1+ \frac{2P \tau \mathbb E\left[ \widetilde g^2 \right] \left(1-H_Q \right)}{\xi^2}} & \mbox{ for Optimal EA},\\ \eta^2 \bigg/ \sqrt{1+ \frac{2P \tau (\mathbb E\left[ \widetilde g \right])^2}{\xi^2 \left(1+R_Q \right)}} & \mbox{ for Equal EA}. \end{cases} \label{def:c} \end{equation} \end{result:var0} In addition to the instantaneous variance and min-max performance limits, one may also be interested in the average variance $\text{Avar}(T)$ as defined by Equation \eqref{def:avar} in Section \ref{sec:prob:form:OU}. The average variance is obtained by integrating \eqref{res:var:theta} over $t\in[0,T]$. Since $\varrho=e^{-t/\tau}$ is the only variable in \eqref{res:var:theta} that depends on $t$, we obtain \begin{align}\begin{split} \text{Avar}(T)&=\frac{\eta^2 \left[1+\eta^2 J \rho' \left\{ 1-\frac{ \mathcal I_T }{ \left( 1-\rho \right)^2} \right\} \right]}{ \sqrt{ \left( \eta^2 J+\rho' \right) \left( \eta^2 J+1/\rho' \right) } }, \\ \mbox{where } \mathcal I_T &\triangleq \frac{1}{T}\int_0^T \left( \varrho -\rho/\varrho \right)^2 \,\mathrm{d} t=\frac{1-\rho^2}{T/\tau}-2\rho. \end{split}\end{align} \begin{figure}[htb] \centering \subfigure[Fixed number of nearest neighbors]{ \includegraphics[width=0.99} \newcommand{\figszb}{0.8 \columnwidth]{collab_asymp_nn_pow.pdf} \label{fig:asymp:nn:pow} } \subfigure[Random geometric graph]{ \includegraphics[width=0.99} \newcommand{\figszb}{0.8 \columnwidth]{collab_asymp_rgg_pow.pdf} \label{fig:asymp:rgg:pow} } \caption{Power-constrained estimation of OU process - Average variance} \vspace{-0.3in} \label{fig:asymp:pow} \end{figure} We use average variance as the performance metric in the following simulation, in which we consider both the aspects discussed in this paper, namely the spatial aggregation procedure and the temporal dynamics. The simulation settings are similar to those considered in Section \ref{sec:snapshot}, which we repeat here for the sake of completeness. The sensor network comprises of $N=10^4$ nodes. We consider $\eta^2=1$, $\xi^2=1$, $\sigma^2=1$, $f_h(h)=\textsf{Rayleigh}(h;1)$ and $f_g(g)=\textsf{Rayleigh}(g;1)$. The exponential drop-off parameter is set to $\tau=1$s and an observation duration of $T_\textsf{obs}=30$s is considered, which is large enough to demonstrate steady state behavior. The observation duration is discretized into $M=1600$ instants for generating the OU process sequence using a first-order autoregressive model. The average variance is obtained as the mean of the deviations from all $M$ estimates. Three values of sampling period are considered for simulation, namely $T=\{0.7,0.4,0.1\}$. The limiting value when $T\rightarrow 0$, $\text{Var}_0$, is also shown on all the graphs. The operating power is chosen as $P=1.4$ to reflect an operating region where substantial performance gain is possible through spatial collaboration. Both NN (Figure \ref{fig:asymp:nn:pow}) and RGG (Figure \ref{fig:asymp:rgg:pow}) topologies are considered to show the applicability of the $Q$-clique results to practical collaboration scenarios. As earlier, both the equal-EA and optimal-EA spatial energy allocation strategies are simulated. The results in Figure \ref{fig:asymp:pow} show that the average variance decreases with $T$. Though we have not rigorously proved that $\text{Avar}(T)$ is monotonically decreasing in $T$, this assertion can be visually verified from Figure \ref{fig:OUproc:var}, by comparing the area under the curves for any two sampling periods ($T=3$ and $T=1.5$, say). This observation coupled with the min-max property of $\text{Var}_0$ leads to the conclusion that an OU process should be sampled as frequently as possible, even if that implies that less energy is available per sampling period (resulting in more noisy samples). However, this assertion is based on the assumption that the sampling noise is temporally white. In practical situations, the sampling errors may become correlated if the samples are obtained too frequently, and caution must be exercised to make sure that the temporal independence assumption is valid. \section{Conclusion} In this paper, we have considered the linear coherent estimation problem in wireless sensor networks and investigated two key aspects. First, we have provided an asymptotic analysis of the single-snapshot estimation problem when the collaboration topology is only partially connected. We achieve this by obtaining the solutions for a family of structured networks and then using those solutions to approximately predict the performance of more sophisticated networks using geometric arguments. Second, we have extended the problem formulation towards the estimation of a time varying signal. In particular, we have derived the instantaneous, average and worst case performance metrics when the signal is modeled as a Gaussian random process with exponential covariance. Both these aspects were investigated under the assumption of spatial and temporal independence among the measurement and channel noise samples. In the future, we plan to relax this assumption and observe the effect of spatial and temporal correlation on the estimation performance. \bibliographystyle{IEEEtran}
\section{Introduction} The topic of indecomposable finite-dimensional representations of the Poincar\'e group was first studied in a systematic way by S. Paneitz (\cite{pan1}\cite{pan2}). In these investigations only representations with one source were considered, though by duality, one representation with 2 sources was implicitly present. The idea of nilpotency was mentioned indirectly in Paneitz's articles, but a more down-to-earth method was chosen there. The results form a part of a major investigation by S. Paneitz and I. E. Segal into physics based on the conformal group. Induction from indecomposable representations plays an important part in this theory. See (\cite{psv}) and references cited therein. The defining representation of the Poincar\'e group, when given as a subgroup of SU(2,2) (see below), is indecomposable. This representation was studied by the present author prior to the articles by Paneitz in connection with a study of special aspects of Dirac operators and positive energy representations of the conformal group (\cite{jak}). Indecomposable representations in theoretical physics have also been used in a major way in a study by G. Cassinelli, G. Olivieri, P. Truini, and V. S. Varadarajan (\cite{raja}). The main object is the Poincar\'e group. In an appendix to the article, the indecomposable representations of the 2-dimensional Euclidean group are considered, and many results are obtained. This group can also be studied by our method, but we will not pursue this here. One small complication is that the circle group is abelian. In the article at hand, we wish to sketch how, by utilizing nilpotency to its fullest extent while using methods from the theory of universal enveloping algebras, a complete description of the indecomposable representations may be reached. In practice, the combinatorics is still formidable, though. It turns out that the method applies to both a class of ordinary Lie algebras and to a similar class of Lie superalgebras. Besides some examples, due to the level of complexity we will only describe a few precise results. One of these is a complete classification of which ideals can occur in the enveloping algebra of the translation subgroup of the Poincar\'e group. Equivalently, this determines all indecomposable representations with a single, 1-dimensional source. Another result is the construction of an infinite-dimensional family of inequivalent representations already in dimension 12. This is much lower than the 24-dimensional representations which were thought to be the lowest possible. The complexity increases considerably, though yet in a manageable fashion, in the supersymmetric setting. Besides a few examples, only a subclass of ideals of the enveloping algebra of the super Poincar\'e algebra will be determined in the present article. \section{Finite-dimensional indecomposable representations of the Poincar\'e group} \label{fd} We are here only interested in what happens on the level of the Lie algebra. Equivalently, we consider a double covering of the Poincar\'e group given by \begin{displaymath} P=\left\{\left(\begin{array}{cc}a&0\\k&{a^\star}^{-1}\end{array}\right) \mid a\in SL(2,{\mathbb R})\; ; k^\star=a^\star k a^{-1} \in gl(2,{\mathbb C})\right\}. \end{displaymath} This is a subgroup of $SU(2,2)$ when the latter is defined by the hermitian form \begin{displaymath} \left(\begin{array}{cc}0&i\\-i&0\end{array}\right) . \end{displaymath} For our purposes, we may equivalently even consider the group \begin{displaymath} P=\left\{\left(\begin{array}{cc}u&0\\z&v\end{array}\right) \mid u,v\in SL(2,{\mathbb C})\; ; z \in gl(2,{\mathbb C})\right\}. \end{displaymath} Let $G_0$ denote the group $SL(2,{\mathbb C})\times SL(2,{\mathbb C})$. Thus, \begin{displaymath} G_0=\left\{\left(\begin{array}{cc}u&0\\0&v\end{array}\right) \mid u,v\in SL(2,{\mathbb C})\;\right\}. \end{displaymath} For what we shall be doing, it does not matter if we work with this group, its Lie algebra, or with $SU(2)\times SU(2)$. It is important to consider the abelian Lie algebra \begin{displaymath} {\mathfrak p}^-=\left\{\left(\begin{array}{cc}0&0\\z&0\end{array}\right) \mid z \in gl(2,{\mathbb C})\right\}. \end{displaymath} along with its enveloping algebra $${\mathcal U}( {\mathfrak p}^-)={\mathcal S}( {\mathfrak p}^-) ={\mathbb C}[z_1,z_2,z_3,z_4].$$ The last equality comes from writing the $2\times 2$ matrix $z$ above as $$z=\left(\begin{array}{cc}z_1&z_2\\z_3&z_4\end{array}\right).$$ In passing we make the important observation that the polynomial $\det z=z_1z_4-z_2z_3$ is invariant in the sense that $\det uzv=\det z$ for all $u,v\in SL(2,{\mathbb C})$. \bigskip We make the basic assumption that all representations and equivalences are over ${\mathbb C}$. This has the powerful consequence that the abelian algebra ${\mathfrak p}^-$ acts nilpotently. The general setting is the following: We consider a reductive Lie algebra ${\mathfrak g}_0$ in which the elements of the abelian ideals are given by semi-simple elements and a nilpotent Lie algebra ${\mathfrak n}$ together with a Lie algebra homomorphism $\alpha$ of ${\mathfrak g}_0$ into the derivations $Der({\mathfrak n})$ of ${\mathfrak n}$. This gives rise, in the usual fashion, to the semi-direct product \begin{displaymath} {\mathfrak g}={\mathfrak g}_0\times_\alpha {\mathfrak n}. \end{displaymath} In this situation a well-known result from algebra (\cite{jac},\cite{hu}) can easily be generalized to include the ${\mathfrak g}_0$ invariance. \medskip Recall that a flag in a vector space $V$ is a sequence of subspaces ${0}=W_0\subsetneq W_1\subsetneq \dots\subsetneq W_r=V$. \begin{Thm}\label{lem1} Suppose given a representation of ${\mathfrak g}$ in some finite-dimensional vector space $V$. Then there is a flag of subspaces such that ${\mathfrak n}$ maps $W_i$ into $W_{i-1}$ for $i=1,\dots,r$ and such that each $W_i$ is invariant and completely decomposable under ${\mathfrak g}_0$. \end{Thm} \medskip We associate a graph to the indecomposable representation $V$ as follows: The nodes are the ${\mathfrak g}_0$ irreducible representations that occur. Two nodes, labeled by irreducibles $V_{1}$ and $V_2$, are connected by an arrow pointing from $V_1$ to $V_2$ if $V_2\subset{\mathfrak n}^-\cdot V_{1}$ inside $V$. If there are multiplicities in the isotypic components the situation becomes more complicated. If the multiplicity at the node $i$ is $n_i$ one can simply place $n_i$ black dots at the node. They can be placed in a stack or in a circle. In case $n_i>1$, $n_j>1$ there may also be a number $a_{i,j}>1$ of arrows from $i$ to $j$, and this needs also to be indicated. The simplest way is just to attach the $n_i$ to each node and to attach the $a_{i,j}$ to the arrow from $i$ to $j$ with the further stipulation that only numbers strictly greater that 1 need to be given. We shall not pursue such details here; see, however, the third of the simple examples below. The theorem above has the immediate consequence that there are no closed paths in this graph. \begin{Rem}The assumption of finite dimension is essential here. Already on the level of the polynomial algebra ${\mathbb C}[z_1,z_2,z_3,z_4]$, if one quotients out by the ideal generated by an inhomogeneous polynomial in $\det z$ there will be closed loops, but the resulting module is infinite-dimensional. If one insists on finite-dimensionality, one must have all homogeneous polynomials in the quotient after a certain degree. Thus $\det z^n$ for some $n$ must be in the ideal. This precludes an inhomogeneous polynomial in $\det z$ since in ${\mathbb C}[T]$ (where $T=\det z$), any inhomogeneous polynomial $p(T)$ is relatively prime to $T^n$ for any $n=0,1,2,\dots$. \end{Rem} Given any such directed graph, any node with arrows only pointing out is as usual called a source. The opposite is called a sink. There is a simple way whereby one may reverse all arrows, thereby turning sources into sinks, and vice versa: Replace $V$ by its dual module $V^\prime$. \medskip Simple situations: \begin{displaymath} \textrm{ One generator } \begin{array}{rcl}&&V_2\\&\nearrow \\V_1\\ &\searrow \\&&V_3\end{array}\qquad\qquad\textrm{ (or its dual...) } \begin{array}{rcl}&&V_2^\prime\\&\swarrow \\V_1^\prime\\ &\nwarrow \\&&V_3^\prime\end{array}\qquad\qquad \end{displaymath} This leads to decomposable representations if the targets (sinks) are equal (respectively if the origins (sources) are equivalent). Otherwise they are indecomposable. \bigskip \begin{displaymath} \textrm{ One source: } \begin{array}{rcl}&&\bullet_{(sink)}\\&\nearrow \\\bullet_{(source)}\\ &\searrow \\&&\bullet_{(sink)}\end{array}\qquad\qquad\textrm{ (or its dual...) } \begin{array}{rcl}&&\bullet_{(source)}\\&\swarrow \\\bullet_{(sink)}\\ &\nwarrow \\&&\bullet_{(source)}\end{array}\qquad\qquad \end{displaymath} \begin{displaymath} \textrm{Two sources, three sinks } \ \begin{array}{rcllc}&&\bullet^2\\&\swarrow &\downarrow&\searrow\\\bullet&&\bullet&&\bullet\end{array}\qquad\qquad\qquad\qquad\qquad \end{displaymath} \bigskip \begin{displaymath} \textrm{ Two generators - two sinks } \begin{array}{rcllc}&&\bullet\\&\nearrow &&\nwarrow\\\bullet&&&&\bullet\\ &\searrow &&\swarrow \\&&\bullet\end{array}\qquad\qquad\qquad\qquad\qquad \end{displaymath} \bigskip \subsection{One generator} We consider only the Poincar\'e algebra. Let $V_0$ denote an irreducible finite-dimensional representation of ${\mathfrak g}_0=sl(2,{\mathbb C})\times sl(2,{\mathbb C})$, given by non-negative integers $(n,m)$ so that the spins are $(\frac{n}2,\frac{m}2)$ and the dimension is $(n+1)(m+1)$. Let $\Pi$ denote an indecomposable finite-dimensional representation of ${\mathcal S}( {\mathfrak p}^-)\times_s {\mathfrak g}_0$ in a space $V_\Pi$, generated by a ${\mathfrak g}_0$ invariant source $V_0$. Let ${\mathcal P}(V_0)$ denote the space of polynomials in the variables $z_1,z_2,z_3,z_4$ with values in $V_0$. This is generated by polynomials of the form $p_0(z_1,z_2,z_3,z_4)\cdot v$ for $v\in V_0$ and $p_0$ a complex polynomial. We consider this a left ${\mathcal S}({\mathfrak p}^-)\times_s {\mathfrak g}_0$ module in the obvious way. The map \begin{equation} p_0(z_1,z_2,z_3,z_4)\cdot v\mapsto \pi(p_0)v \end{equation} is clearly ${\mathcal S}( {\mathfrak p}^-)\times_s {\mathfrak g}_0$ equivariant. The decomposition of the ${\mathfrak g}_0$ module ${\mathcal S}( {\mathfrak p}^-)$ is well-known and is given by the representations $\textrm{spin}(\frac{n}2,\frac{n}2)$ for $n=0,1,2,\dots$. Each occurs with infinite multiplicity due to the invariance of $\det z$ under ${\mathfrak g}_0$. We will describe these representations in detail below. \medskip The decomposition of ${\mathcal P}(V_0)$ into irreducible ${\mathfrak g}_0$ representations follows easily from this using the well-known decomposition of the tensor product of two irreducible representations of $su(2)$. The decomposition of ${\mathcal P}(V_0)$ is in general more degenerate than what results from the invariance of $\det z$. It is clear that there exists a sub-module ${\mathcal I}\subseteq {\mathcal P}(V_0)$ such that \begin{displaymath} {\mathcal P}(V_0)/{\mathcal I}\equiv V_\Pi. \end{displaymath} The finite-dimensionality assumption on $V_\Pi$ then implies that ${\mathcal I}$ contains all homogeneous polynomials of a degree greater than or equal to some fixed degree, say $d_0$. Since there are only a finite number of linearly independent homogeneous polynomials in ${\mathcal P}(V_0)$ of degree $d_0$, it follows that there exists a finite number of elements $p_1,p_2,\dots,p_j$ in ${\mathcal P}(V_0)$ (these may be chosen for instance as highest weight vectors) such that if ${\mathcal I}\langle p_1,p_2,\dots, p_j\rangle$ denotes the ${\mathcal S}( {\mathfrak p}^-)\times_s {\mathfrak g}_0$ submodule generated by these elements, then \begin{displaymath} V_\Pi\equiv {\mathcal P}(V_0)/{\mathcal I}\langle p_1,p_2,\dots, p_j\rangle. \end{displaymath} We assume that the number $j$ of polynomials is minimal. \medskip Once the elements $p_1,p_2,\dots,p_j$ are known, it is possible to construct the whole graph as above. In particular, {\bf the sinks in $V_\Pi$ can be directly determined from this}. See Proposition~\ref{clearprop} below for a simple example that indicates how. In case $\dim V_0$ is large the task, of course, will be more cumbersome. \medskip \begin{Ex} As is well known, we have that $${\mathfrak p}^-\otimes \textrm{spin}(\frac12,0)=\textrm{spin}(1,\frac12)\oplus \textrm{spin}(0,\frac12).$$ If we mod out by all second order polynomials, and possibly one of the first order polynomial representations, we get the following 3 indecomposable representations: \begin{itemize} \item $\textrm{spin}(\frac12,0)\rightarrow \textrm{spin}(0,\frac12)$. This 4-dimensional representation comes from the the defining representation. \item $\textrm{spin}(\frac12,0)\rightarrow \textrm{spin}(1,\frac12)$. This is an 8-dimensional representation. \item $\textrm{spin}(\frac12,0)\rightarrow \textrm{spin}(0,\frac12), \textrm{spin}(1,\frac12)$. This is a 10-dimensional representation which includes the two former. \end{itemize} \medskip Proceeding analogously, $${\mathfrak p}^-\otimes \textrm{spin}(1,0)=\textrm{spin}(\frac32,\frac12)\oplus \textrm{spin}(\frac12,\frac12)$$ leads to inequivalent representations in dimensions 7,11,15. Similarly, $${\mathfrak p}^-\otimes \textrm{spin}(\frac12,\frac12)=\textrm{spin}(0,0)\oplus \textrm{spin}(1,0)\oplus \textrm{spin}(0,1)\oplus \textrm{spin}(1,1)$$ leads to indecomposable representations in dimensions 5, 7, 8, 10, 11, 13, 16, 17, 19, 20. Several dimensions here carry a number of inequivalent representations. Together with duals of these or versions obtained as mirror images by interchanging the spins, these exhaust all the known representations in dimensions less than or equal to 8 with the exception of a 6-dimensional representation which we describe in Example~\ref{ex2}. \end{Ex} \medskip \subsection{Special case: Ideals in ${\mathcal U}({\mathfrak p}^-)$} It is well-known that there is a decomposition \begin{displaymath} {\mathcal U}({\mathfrak p}^-)=\oplus W_{r,s} \end{displaymath} into ${\mathfrak g}_0$ representations, where the subspace $W_{r,s}$ may be defined through its highest weight vector, say $z_1^r\det z^s$. This is possible since each representation occurs with multiplicity one. Denote this representation simply by $[r,s]$. It is elementary to see that the action of ${\mathfrak p}^-$ on the left on ${\mathcal U}({\mathfrak p}^-)$, when expressed in terms of representations, is given as follows. All arrows represent non-trivial maps. \begin{equation} \begin{array}{cccccccc}\\\downarrow\\\left[1,0\right]\\\downarrow&\searrow\\\left[2,0\right]&&\left[0,1\right]\\\downarrow&\searrow&\downarrow\\\left[3,0\right]&&\left[1,1\right]\\\downarrow&\searrow&\downarrow&\searrow\\\left[4,0\right]&&\left[2,1\right]&&\left[0,2\right]\\\downarrow&\searrow&\downarrow&\searrow&\downarrow\\\left[5,0\right]&&\left[3,1\right]&&\left[1,2\right]\\\downarrow&\searrow&\downarrow&\searrow&\downarrow&\searrow\\\left[6,0\right]&&\left[4,1\right]&&\left[2,2\right]&&\left[0,3\right]\\\vdots&&\vdots&&\vdots&&\vdots \end{array} \label{tree} \end{equation} Any ideal ${\mathcal I}\subseteq {\mathcal U}({\mathfrak p}^-)$ that has finite codimension must clearly contain some $z_1^{r_1}$ for some minimal $r_1\in {\mathbb N}$ (we omit the trivial case of codimension 0). Since we are assuming that the ideals are ${\mathfrak g}_0$ invariant, if some other element $z_1^{r_2}p(\det z)$ is in the ideal then first of all we can assume $r_2<r_1$ and secondly we can assume that the polynomial $p$ is homogeneous; $p(\det z)=\det z^{s_2}$ for some $s_2>0$. The latter inequality follows by the minimality of $r_1$. Thus the following is clear: \begin{Prop} \label{clearprop} Any ${\mathfrak g}_0$ invariant ideal ${\mathcal I}\subseteq {\mathcal U}({\mathfrak p}^-)$ of finite codimension is of the form \begin{equation} {\mathcal I}={\mathfrak g}_0\cdot \langle z_1^{r_1}\det z^{s_1},z_1^{r_2}\det z^{s_2},\cdots, z_1^{r_t}\det z^{s_t}\rangle \label{eqw} \end{equation} for some positive integers $r_1>r_2>\dots>r_t$ and integers $0=s_1<s_2<\cdots<s_t$. If the set is minimal, then furthermore $$\forall j=2,3,\dots, t: s_1+s_2+\dots + s_j\leq r_1-r_j.$$ Any set of such integers determine an invariant ideal of finite codimension. The sinks in the quotient module are $$z_1^{r_1-s_2}\det z^{s_2-1},z_1^{r_1-s_2-s_3}\det z^{s_3-1}\dots z_1^{r_1-s_2-\dots-s_t}\det z^{s_t-1},\textrm{ and } \det z^{s_t+r_t-1}.$$ \end{Prop} \medskip \begin{Ex}\label{ex2} If we mod out by all second order polynomials, that is by $z_1^2$ and $\det z$, we get the 5-dimensional indecomposable representation $$\textrm{spin}(0,0)\rightarrow \textrm{spin}(\frac12,\frac12).$$ If we instead mod out by $z_1^2$ and $z_1\det z$ we get the 6-dimensional indecomposable representation $$\textrm{spin}(0,0)\rightarrow \textrm{spin}(\frac12,\frac12)\rightarrow \textrm{spin}(0,0).$$ \end{Ex} \medskip \begin{Ex} The representations determined by ideals are easily written down, though some finer details from the representation theory of $su(2)$ will have to be invoked to get the precise form. In simple examples like the last in the previous example, everything follows immediately since there is no need to be precise about the relative scales in the 3 spaces: \begin{eqnarray} {\mathfrak p}^-\ni \underline{p}=(p_1,p_2,p_3,p_4)\mapsto \left(\begin{array}{cccccc}0&0&0&0&0&0\\p_1&0&0&0&0&0\\p_2&0&0&0&0&0\\p_3&0&0&0&0&0\\p_4&0&0&0&0&0\\0&p_4&-p_3&-p_2&p_1&0 \end{array}\right). \end{eqnarray} An element $(u,v)$ in the diagonal subgroup $G_0$ (see Section~\ref{fd}) acts as $0\oplus u\otimes (v^t)^{-1}\oplus 0$. Notice that the matrix with just $p_1$ corresponds to a map which sends the constant 1 to the polynomial $p_1z_1$ and sends the polynomial $z_4$ to $p_1\det z$ and all other first order polynomials $z_1,z_2,z_3$ to 0. \end{Ex} \medskip \subsection{Two sources and 2 sinks} We here consider the Poincar\'e algebra. Consider the situation previously depicted under `Simple situations' where there is one source and two sinks. The resulting representations may be written as \begin{equation}\label{repg1} \left(\begin{array}{cc}0&0\\w&0\end{array}\right)\mapsto \left(\begin{array}{ccc}0&0&0\\F(w)&0&0\\G(w)&0&0 \end{array}\right)\; , \left(\begin{array}{cc}u&0\\0&v\end{array}\right)\mapsto \left(\begin{array}{ccc}\tau_1(u,v)&0&0\\0&\tau_2(u,v)&0\\0&0&\tau_3(u,v) \end{array}\right). \end{equation} With this one can easily write down a representation with 2 sources and 2 sinks, indeed a 4-parameter family given by elements $(\alpha,\beta,\gamma,\delta)\in {\mathbb C}^4$: \begin{equation}\label{repg2} \left(\begin{array}{cccc}0&0&0&0\\0&0&0&0\\\alpha\cdot F(w)&\beta\cdot F(w)&0&0\\\gamma \cdot G(w)&\delta\cdot G(w)&0&0 \end{array}\right)\textrm{ resp. } \left(\begin{array}{cccc}\tau_1(u,v)&0&0&0\\0&\tau_1(u,v)&0&0\\0&0&\tau_2(u,v)&0\\0&0&0&\tau_3(u,v) \end{array}\right). \end{equation} \medskip In this case, there is a continuum of inequivalent representations and they are generically indecomposable. This lead to a continuum already in dimension 12 where the two sources are equal and 2-dimensional - say spin($\frac12$,0), and the two sinks are spin($1$,$\frac12$) and spin(0,$\frac12$) or, also in dimension 12, the 2 sources are the 4-dimensional spin($\frac12$,$\frac12$), and the sinks are spin($1$,0) and spin(0,0), or in dimension 16 where one is the 2-dimensional spin ($\frac12$,0) and the other is the 6-dimensional spin($\frac12$,$1$) and the targets are spin(0,$\frac12$) and spin(1,$\frac12$). The moduli space in these cases is ${\mathbb C}{\mathbb P}^1$. Specifically, the indecomposable is determined by a point $(p,q) \in {\mathbb C}{\mathbb P}^1\times {\mathbb C}{\mathbb P}^1$ giving relative scales on the arrows. Here, $p\equiv(\alpha,\beta)$ and $q\equiv(\gamma,\delta)$ in the above representation. Two such points $(p_1,q_1)$ and $(p_2,q_2)$, are equivalent if there is an element $g\in GL(2,{\mathbb C})$ such that $(p_2,q_2)=(gp_1,gp_2)$. \bigskip \section{Supersymmetry} The previous considerations are now extended to the supersymmetric setting as follows: Let $H_1, H_2$, and $H_3$ be reductive matrix Lie groups with Lie algebras ${\mathfrak h}_1, {\mathfrak h}_2$, and ${\mathfrak h}_3$, respectively. We assume that possible abelian ideals are represented by semi-simple elements. We consider an irreducible representation of each of these Lie algebras; $V_1,V_2$, and $V_3$. We identify the representation with the space in which it acts. We denote furthermore the dual representation of a representation $V$ by $V^\prime$ (this is the ${\mathbb C}$ linear dual). Let \begin{eqnarray}W_1= \hom(V_1,V_2)\equiv V_1^\prime\otimes V_2 &;&W_2= \hom(V_2,V_3)\equiv V_2^\prime\otimes V_3\\\qquad \textrm{and} \qquad Z= \hom(V_1,V_3)\equiv V_1^\prime\otimes V_3. \end{eqnarray} The Lie superalgebra ${\mathfrak g}_{super}={\mathfrak g}_{super}({\mathfrak h}_1, {\mathfrak h}_2,{\mathfrak h}_3,V_1,V_2,V_3)$ is defined as \begin{eqnarray} &{\mathfrak g}_{super}=\\&\left\{\left(\begin{array}{ccc}a&0&0\\w_1&g&0\\z&w_2&b\end{array}\right) \mid a\in{\mathfrak h}_1\; , g \in {\mathfrak h}_2\; , b\in {\mathfrak h}_3\; , w_1\in W_1\;,w_2\in W_2\;, \textrm{ and }z\in Z \right\}.\nonumber \end{eqnarray}The odd part is given as \begin{eqnarray} {\mathfrak g}_{super}^1=\left\{\left(\begin{array}{ccc}0&0&0\\w_1&0&0\\0&w_2&0\end{array}\right) \mid w_1\in W_1\;,\textrm{ and }w_2\in W_2\;\right\}.\nonumber \end{eqnarray} Let \begin{eqnarray} {\mathfrak n}_{super}=\left\{\left(\begin{array}{ccc}0&0&0\\w_1&0&0\\z&w_2&0\end{array}\right) \mid w_1\in W_1\;,w_2\in W_2\; ,\textrm{ and }z\in Z\right\}.\nonumber \end{eqnarray} and \begin{eqnarray} {\mathfrak g}^r_{super}=\left\{\left(\begin{array}{ccc}a&0&0\\0&g&0\\0&0&b\end{array}\right) \mid a\in{\mathfrak h}_1\; , g \in {\mathfrak h}_2\; , \textrm{ and }b\in {\mathfrak h}_3\; \right\}.\nonumber \end{eqnarray} Obviously, ${\mathfrak n}_{super}$ is a maximal nilpotent ideal and ${\mathfrak g}_{super}^r$ is the reductive part. We let \begin{eqnarray} {\mathfrak p}^-=\left\{\left(\begin{array}{ccc}0&0&0\\0&0&0\\z&0&0\end{array}\right) \mid z\in Z\right\}.\nonumber \end{eqnarray} Then ${\mathfrak g}_{super}^0={\mathfrak g}^r_{super}\oplus {\mathfrak p}^-$. We then have the following super algebraic generalization of Theorem~\ref{lem1}: \begin{Thm}Consider a finite-dimensional representation $V_{super}$ of ${\mathfrak g}_{super}$. Then there is a flag of subspaces $\{0\}=W_0\subsetneq W_1\subsetneq\cdots\subsetneq W_{r-1}\subsetneq W_r=V_{super}$ such that each $W_i$ is invariant and completely reducible under ${\mathfrak g}_{super}^0$ and such that ${\mathfrak n}_{super}$ maps $W_i$ into $W_{i-1}$ for $i=1,\dots,r$. \end{Thm} Thus, the previous treatment with directed graphs and ideals carry over. Naturally, the picture gets more complicated with the odd generators. The most simple thing to consider would be the ${\mathfrak g}_{super}$ module ${\mathcal U}({\mathfrak n}_{super})$, but even here the situation is complex, though in principle tractable. Consider as an example the case of the simplest super Poincar\'e algebra, \begin{eqnarray} &{\mathfrak g}_{super}^{P}=\\&\left\{\left(\begin{array}{ccc}a&0&0\\w_1&0&0\\z&w_2&b\end{array}\right) \mid a,b\in sl(2,{\mathbb C}) , w_1 \in M(1,2), w_2\in M(2,1)\;, \textrm{ and }z\in M(2,2) \right\}.\nonumber \end{eqnarray} \medskip \medskip Let $W_1=M(1,2)$ and $W_2=M(2,1)$. We number the spaces $$\begin{array}{ccc}1&W_2&W_2\wedge W_2\\W_1&W_1\wedge W_2&W_1\wedge W_2\wedge W_2\\W_1\wedge W_1&W_1\wedge W_1\wedge W_2&W_1\wedge W_1\wedge W_2\wedge W_2\end{array}=\begin{array}{ccc}1&2&3\\4&5&6\\7&8&9\end{array}.$$ We then have that \begin{equation} {\mathcal U}({\mathfrak n}_{super})=\sum_{i=1}^9 {\mathcal U}({\mathfrak p}^-)\otimes {\mathcal U}({\mathfrak q}_{super}^1)_i. \label{nsu} \end{equation} \medskip Each of the 9 summands is invariant under ${\mathfrak g}^r_{super}$. The representations corresponding to this are given right below . Here, $n,d$ are independent non-negative integers (in a few obvious cases, n must furthermore be non-zero). The labels $\uparrow$ and $\downarrow$ may be taken just as part of a short hand notation that are defined by the stated equations. They are listed here for convenience even though they are not used directly. One can ascertain useful information from them about how the various pieces occur in the tensor products of $su(2)$ representations. \begin{eqnarray*} 1\left[n,n,d\right]&\oplus&\\ 2\uparrow\left[n,d\right]=2\left[n-1,n,d\right]&\oplus&2\downarrow\left[n,d\right]=2\left[n+1,n,d\right]\\ 3\left[n,n,d\right]&\oplus&\\ 4\uparrow\left[n,d\right]=4\left[n,n-1,d\right]&\oplus&4\downarrow\left[n,d\right]=4\left[n,n+1,d\right]\\ 5\uparrow\uparrow\left[n,d\right]=5\left[n-1,n-1,d\right]&\oplus&5\uparrow\downarrow\left[n,d\right]=5\left[n-1,n+1,d\right]\\ 5\downarrow\uparrow\left[n,d\right]=5\left[n+1,n-1,d\right]&\oplus&5\downarrow\downarrow\left[n,d\right]= 5\left[n+1,n+1,d\right]\\ 6\uparrow\left[n,d\right]=6\left[n,n-1,d\right]&\oplus&6\downarrow\left[n,d\right]=6\left[n+1,n,d\right]\\ 7\left[n,n,d\right]&\oplus&\\ 8\uparrow\left[n,d\right]=6\left[n-1,n,d\right]&\oplus&8\downarrow\left[n,d\right]=8\left[n+1,n,d\right]\\ &\oplus&9\left[n,n,d\right] \end{eqnarray*} A further complication is that there are representations in different spaces that are equivalent under ${\mathfrak g}^r_{super}$: \begin{eqnarray*} \label{relations} 8\uparrow\left[n,d\right]&\leftrightarrow&4\uparrow\left[n-1,d+1\right]\\ 8\downarrow\left[n,d\right]&\leftrightarrow&4\left[n+1,d\right]\\ 5\downarrow\downarrow\left[n,d\right]&\leftrightarrow&1\left[n+1,d\right]\\ 5\uparrow\uparrow\left[n,d\right] &\leftrightarrow&1\left[n-1,d+1\right]\\ 5\downarrow\downarrow\left[n-1,d+1\right],5\uparrow\uparrow\left[n+1,d\right]&\leftrightarrow&9\left[n,d\right]\\ 5\downarrow\downarrow\left[n-1,d+1\right],5\downarrow\downarrow\left[n+1,d\right]&\leftrightarrow&9\left[n,d\right]\\ 6\uparrow\left[n,d\right]&\leftrightarrow&2\downarrow\left[n-1,d+1\right]\\ 6\uparrow\left[n,d\right]&\leftrightarrow&2\downarrow\left[n-1,d+1\right]\\ 6\downarrow\left[n,d\right]&\leftrightarrow&2\uparrow\left[n+1,d\right]\\ 6\downarrow\left[n,d\right]&\leftrightarrow&2\uparrow\left[n+1,d\right] \end{eqnarray*} To each finite-dimensional representation $V_r$ of ${\mathfrak g}^r_{super}$ (may be reducible), the general object of interest is the left module \begin{equation} {\mathcal U}({\mathfrak n}_{super})\cdot V_r=\sum_{i=1}^n {\mathcal U}({\mathfrak p}^-)\otimes {\mathcal U}({\mathfrak q}_{super}^1)_i\cdot V_r. \label{nsu1} \end{equation} To further analyze this we have to choose a PBW-type basis. We will do this in the indicated fashion with ${\mathcal U}({\mathfrak p}^-)$ to the left and with furthermore $W_1$ always to the left of $W_2$. \begin{Ex}Assume that ${\mathcal U}({\mathfrak p}^-)$ acts trivially on the space $V_r$. The resulting module is then \begin{equation}\bigwedge({\mathfrak g}^1_{super})\cdot V_r, \label{triv} \end{equation} or some of the subrepresentations thereof. The exterior algebra $\bigwedge({\mathfrak g}^1_{super})$ occurs because the $W_1$, $W_2$ anticommute in the considered quotient. \end{Ex} \medskip Observe that in the sum (\ref{nsu1}) the summand \begin{equation} {\mathcal U}_{2,3,5,6,8,9}({\mathfrak n}_{super})\cdot V_r=\sum_{i=2,3,5,6,8,9} {\mathcal U}({\mathfrak p}^-)\otimes {\mathcal U}({\mathfrak q}_{super}^1)_i\cdot V_r \label{nsu2} \end{equation} is invariant. We may then pass to a general subclass of indecomposable modules by first taking the quotient by this. The vector space that results is \begin{equation} {\mathcal U}_{rest}({\mathfrak n}_{super})\cdot V_r=\sum_{i=1,4,7} {\mathcal U}({\mathfrak p}^-)\otimes {\mathcal U}({\mathfrak q}_{super}^1)_i\cdot V_r. \label{nsu3} \end{equation} If we let ${\mathcal U}({\mathfrak p}^-)_{\geq s}$ be the ideal generated by all homogeneous elements of degree $s$, it is easy to see that for each $s=0,1,2,\dots$, the space \begin{equation} {\mathcal U}_{rest}^s({\mathfrak n}_{super})\cdot V_r= {\mathcal U}({\mathfrak p}^-)_{s+2}\cdot V_r \oplus {\mathcal U}({\mathfrak p}^-)_{s+1}\cdot W_1\cdot V_r\oplus {\mathcal U}({\mathfrak p}^-)_{s}\cdot (W_1\wedge W_1)\cdot V_r \label{nsu4} \end{equation} is invariant. \begin{Ex} Let $V_r^1$ be the irreducible 2-dimensional representation which is only non-trivial on the ${\mathfrak h}_1$ piece of the reductive part. The defining representation of ${\mathfrak g}_{super}^{P}$ is a subrepresentation of the quotient $${\mathcal U}_{rest}({\mathfrak n}_{super})\cdot V^1_r/{\mathcal U}_{rest}^0({\mathfrak n}_{super})\cdot V^1_r.$$ Indeed, we just have to limit ourselves further by removing two appropriate ${\mathfrak g}_{super}^r$ representations. \end{Ex} Returning to the more general situation, let us assume from now on that $V_r$ is the trivial 1-dimensional module. We are thus left with the space \begin{equation} {\mathcal U}_{rest}({\mathfrak n}_{super})= {\mathcal U}({\mathfrak p}^-) \oplus {\mathcal U}({\mathfrak p}^-)\cdot W_1\oplus {\mathcal U}({\mathfrak p}^-)\cdot (W_1\wedge W_1). \label{nsu5} \end{equation} The general form of the representation is (we give only the $W_1,W_2$ operators) \begin{equation} \left(\begin{array}{cccc}0&(w_2^1p_{11}+w_2^2p_{21})1_{\mathcal P}&(w_2^1p_{12}+w_2^2p_{22})1_{\mathcal P}&0\\w_1^11_{\mathcal P}&0&0&-(w_2^1p_{12}+w_2^2p_{22})1_{\mathcal P}\\w_1^21_{\mathcal P}&0&0&(w_2^1p_{11}+w_2^2p_{21})1_{\mathcal P}\\0&-w_1^21_{\mathcal P}&w_1^11_{\mathcal P}&0 \end{array}\right) . \label{genfo} \end{equation} Here, each block corresponds to a space of polynomials. The operators $p_{i,j}$ are multiplication operators in ${\mathcal U}({\mathfrak p}^-)$ - and hence also such operators in the given space. Notice that they increase the degree of the target by 1. The symbol $1_{\mathcal P}$ denotes the identity operator. \medskip The finer details are given as follows, where the arrows point upwards come from $W_2$ and those pointing downwards come from $W_1$. \begin{equation} \begin{array}{cccccccccccccccc}&&1\left[n,d+1\right]&&\\&\nearrow&&\nwarrow&\\4\downarrow\left[n-1,d+1\right]&&&&4\uparrow\left[n+1,d\right]\\&\nwarrow&&\nearrow&\\ &&7\left[n,d\right]&& \end{array} \end{equation} \begin{equation} \begin{array}{cccccccccccccccc}&&1\left[n,d\right]&\\&\swarrow&&\searrow\\4\downarrow\left[n,d\right]&&&&4\uparrow\left[n,d\right] \\&\nwarrow&&\nearrow\\ &&7\left[n,d\right] \end{array} \end{equation} \medskip Any invariant ideal ${\mathcal I}_{super}$ in ${\mathcal U}_{rest}({\mathfrak n}_{super})$ contains a sum of the form \begin{equation} {\mathcal I}_1({\mathfrak p}^-) \oplus {\mathcal I}_4({\mathfrak p}^-)\cdot W_1\oplus {\mathcal I}_7({\mathfrak p}^-)\cdot (W_1\wedge W_1), \label{nsu6} \end{equation} where ${\mathcal I}_1({\mathfrak p}^-)\subseteq {\mathcal I}_4({\mathfrak p}^-)\subseteq {\mathcal I}_7({\mathfrak p}^-)\subseteq {\mathcal U}({\mathfrak p}^-)$ are ${\mathfrak p}^-$ ideals. These are precisely the ideals determined in Section~1. \begin{Prop} If we have a representation $7\left[n,d\right]\in {\mathcal I}_7({\mathfrak p}^-)$ then $4\left[n,d+1\right]\in {\mathcal I}_4({\mathfrak p}^-)$ and $1\left[n,d+1\right]\in {\mathcal I}_1({\mathfrak p}^-)$. Furthermore, $4\uparrow\left[n,d\right], 4\downarrow\left[n,d\right]\in {\mathcal I}_{super}$. In particular, $${\mathfrak p}^-\cdot {\mathfrak p}^-\cdot{\mathcal I}_7({\mathfrak p}^-)\subseteq {\mathcal I}_1.$$ \end{Prop} This result in principle solves the problem but there are still extremely many cases - even if we start with an ideal ${\mathcal I}_7$ and ask for how many configurations of the ideals ${\mathcal I}_1, {\mathcal I}_4$ that are possible. We refrain from pursuing this further and just give a low-dimensional example. \begin{Ex}Let ${\mathcal I}_7={\mathcal I}\langle z_1\rangle$. The following list is exhaustive and each case occurs. \begin{itemize} \item ${\mathcal I}_1={\mathcal I}\langle z_1\rangle$ then ${\mathcal I}_4={\mathcal I}\langle z_1\rangle$. \item ${\mathcal I}_1={\mathcal I}\langle z^2_1\rangle$ then either ${\mathcal I}_4={\mathcal I}\langle z^2_1\rangle$, ${\mathcal I}_4={\mathcal I}\langle z^2_1, \det z\rangle$, or ${\mathcal I}_4={\mathcal I}_7$. \item ${\mathcal I}_1={\mathcal I}\langle z^2_1, \det z\rangle$ then ${\mathcal I}_4={\mathcal I}\langle z^2_1, \det z\rangle$. \item ${\mathcal I}_1={\mathcal I}\langle z^3_1, \det z\rangle$ then ${\mathcal I}_4={\mathcal I}\langle z^2_1, \det z\rangle$. \item ${\mathcal I}_1={\mathcal I}\langle z^3_1, z_1\det z\rangle$ then ${\mathcal I}_4={\mathcal I}\langle z^2_1, \det z\rangle$ or ${\mathcal I}_4={\mathcal I}^2_1$. \end{itemize} \end{Ex} \medskip
\section{Introduction} The discovery of a Higgs-like boson at the LHC raises several questions, one of them being an understanding of the origin of the associated mechanism of Spontaneous Symmetry Breaking (SSB). One of the most exciting explanations is a possible connection of the mechanism to the existence of an extra dimension \cite{Hoso}. In this, "Gauge-Higgs Unification" (GHU) scenario, the Higgs boson comes from some of the extra dimensional components of the five-dimensional gauge field. The model we will discuss here is an $SU(2)$ gauge theory in five dimensions with the fifth dimension compactified on the $S^1/Z_2$ orbifold. The embedding of the orbifold action in the gauge field $A_M^A$ where $M=\mu, 5$ is a Lorentz index and $A=1,2,3$ is a gauge index, is such that $A_\mu^3$ with $\mu=0,1,2,3$ and $A_5^{1,2}$ are even, and all other components are odd. The latter is the "Higgs", a complex scalar in the fundamental representation of the unbroken $U(1)$ boundary symmetry. This is the simplest, prototype model. Its generalization to an $SU(3)$ bulk symmetry leaves an $SU(2)\times U(1)$ symmetry on the boundary with two complex doublet Higgs fields, in the fundamental representation of the $SU(2)$ factor. We believe that if the mechanism of SSB is present in the simpler system, it will be generically present also in the more complicated cases, so we study it first in the $SU(2)$ model. We also take the point of view that SSB is driven by pure gauge dynamics, a fact that has been observed in earlier Monte Carlo simulations \cite{MC1}. In fact, a perturbative 1-loop computation of the Coleman-Weinberg potential does not yield SSB in the pure gauge system at infinite cut-off \cite{Kubo}. With a finite cut-off it is possible in principle to have SSB \cite{NFM}, however it is not possible to prove this in the perturbative context since the quantum theory is non-renormalizable. All this points to the necessity for developing a non-perturbative analytical tool which can probe the system near its bulk phase transition \cite{Creutz}, where a scaling regime with suppressed cut-off effects might exist. Such a formalism has been developed in \cite{MF1,MF2} and it is an expansion in fluctuations around a Mean-Field (MF) background \cite{DZ}. There is serious evidence that the MF expansion describes the non-perturbative system with periodic boundary conditions faithfully \cite{FrancAM}, so we will now apply it to the orbifold model as well. The parameters of the model are the dimensionless five-dimensional lattice coupling $\beta$, the anisotropy $\gamma$ and the lattice size which is set by $(T,L)$ points along the $(\mu=0, \mu=1,2,3)$ directions and $N_5+1$ points along the extra dimension. $L$ will be taken always large enough so that physics does not depend on it. Typically this happens when $L\ge 200$ approximately. As the lattice action, the Wilson plaquette action is used, with the orbifold boundary conditions and the anisotropy appropriately implemented in it \cite{MF2}. \section{The Mean-Field expansion} The first step in the MF formalism is to trade the gauge links $U$ of the lattice with unconstrained complex variables $v$ and a set of Lagrange multipliers $h$ that ensure that the memory of the gauge nature of the links is not lost. Then, the gauge links can be integrated out. The resulting effective action is then minimized with respect to the left over degrees of freedom $v$ and $h$. The saddle point solution defines the MF background. We will be considering lattices wich are isotropic in the $\mu$ directions and have an anisotropy $\gamma$ along the fifth dimension. Consequently, the background is $\overline v_0(n_5)$ along four-dimensional hyperplanes and $\overline v_{05}(n_5+1/2)$ along the fifth dimension. Here $n_5$ is the discrete label of extra dimensional points on the lattice. The phases of the system are defined as follows: \begin{itemize} \item Confined phase: $\overline v_0(n_5), \overline v_{05}(n_5+1/2) = 0$. \item Layered phase: $\overline v_0(n_5) \ne 0$, $\overline v_{05}(n_5+1/2) = 0$. \item Deconfined phase: $\overline v_0(n_5), \overline v_{05}(n_5+1/2) \ne 0$. \end{itemize} The above solutions are found by solving numerically the non-linear algebraic equations that reflect the saddle point of the path integral. The boundary conditions ensure that this anisotropic solution is not an artifact, since translational invariance is broken along the fifth dimension by the presence of the boundaries. We parametrize the fluctuations around the MF background as \begin{eqnarray} v&=&v_0+iv_A\sigma^A, v_{0,A}\in C \nonumber\\ h&=&h_0+ih_A\sigma^A, h_{0,A}\in C \, .\nonumber \end{eqnarray} These are introduced in the formalism by the substitution \begin{eqnarray} && v \longrightarrow \overline v_0 + v \nonumber\\ && h \longrightarrow \bar h_0 + h \nonumber \end{eqnarray} in the path integral and performing a derivative expansion on the effective action and on the gauge invariant observable $\cal O$. The second derivative part of this expansion defines the lattice propagator \begin{equation} {K}_{M'M''}^{-1} = {K}_{M'M''}^{-1}(p',n'_5,M',\alpha' ;p'',n''_5,M'',\alpha'')\, , \end{equation} with $p=(p_0,p_k)$ the four-dimensional lattice momenta and $\alpha=0,A$ a gauge index. The quantity \begin{equation} \langle {\cal O} \rangle = {\cal O}[{\overline v_0}] + \frac{1}{2}{\rm tr} \left\{\frac{\delta^2{\cal O}}{\delta v^2}\Biggr|_{\overline v_0} K^{-1}\right\} \label{correction} \end{equation} defines an observable to first order in the fluctuations. The time-dependent correlator $C(t)$ is then defined as \begin{equation} C (t) = \langle {\cal O} (t_0+t){\cal O} (t_0)\rangle - \langle {\cal O} (t_0+t)\rangle \langle {\cal O} (t_0)\rangle\, , \end{equation} with $t_0$ an arbitrary initial time and the mass of the associated ground state is extracted from \begin{equation} m = \lim_{t\to \infty} \ln \frac{C (t)}{C (t-1)}\, .\label{mass} \end{equation} The main observable of our interest will be the Wilson Loop ${\cal O}_W(r,t)$ of length $r$ along one of the dimensions on the boundary. The static potential extracted as \begin{equation} t\to \infty : \hskip .5cm {\rm e}^{-V(r) t} \simeq \; <{\cal O}_W> \end{equation} contains the key information about SSB: if the static potential is of a four-dimensional Yukawa form with the Yukawa mass identified with the mass of the boundary gauge boson, then we have a spontaneously broken $U(1)$ on the boundary. In this case, we are entitled to call this gauge boson a "$Z$ boson". To leading order in the MF expansion, the two types of gauge boson exchange diagrams appearing on fig. \ref{Wilson} dominate the Wilson Loop. \begin{figure}\centering \resizebox{6cm}{!}{\includegraphics[angle=0]{WilsonOrb.eps}} \caption{Contributions to the static potential on the boundary of the orbifold: gauge boson exchange and self energy.} \label{Wilson} \end{figure} Specializing to our $SU(2)$ model, for the static potential, we arrive at the result \cite{MF2}: \begin{eqnarray} &&V(r)=-\log({\overline v_0(0)^2})-\frac{1}{2}\frac{1}{L^3} \frac{1}{(\overline v_0(0))^2}\sum_{p_k'}\nonumber\\ && \Biggl\{\frac{1}{3}\sum_k\Bigl[ 2\cos{(p_k'r)}+2\Bigr] {K}^{-1} \left((0,p_k'),0,0,0;(0,p_k'),0,0,0\right)\nonumber\\ &+& \frac{1}{3}\sum_k\Bigl[ 2\cos{(p_k'r)}-2\Bigr] {K}^{-1}\left((0,p_k'),0,0,3;(0,p_k'),0,0,3\right)\Biggr\}\, . \label{potb} \end{eqnarray} There is a similar expression for the Wilson Loop parallel to the above, sitting in the middle of the fifth dimension. The other observable that we will use is the one with scalar quantum numbers, corresponding to the Higgs field. It is basically a pair of Polyakov Loops (projected on the orbifold) along the fifth dimension separated in the time direction, exchanging a gauge boson. The computation of this diagram yields the Higgs correlator $C_H(t)$ (for details see \cite{MF2}): \begin{equation} C_H(t) = \frac{8}{L^3 T}(P_0^{(0)})^2 \Pi^{(1)}_{\langle 1,1\rangle}(0,0) \,, \end{equation} where $P_0^{(0)}$ is the Polyakov loop evaluated on the background and $\Pi^{(1)}_{\langle 1,1\rangle}(0,0)$ is \newpage \begin{eqnarray} \Pi^{(1)}_{\langle 1,1\rangle}(0,0) &=& 2\sum_{p_0'}\cos{p_0't} \sum_{n_5',n_5''} \sum_{r=0}^{N_5-1}\frac{\delta_{n_5', r}}{\overline v_0(r+1/2)}\cdot \nonumber\\ && {K}^{(-1)}((p_0',{\vec 0}),n_5',5,0;(p_0',{\vec 0}),n_5'',5,0) \sum_{r=0}^{N_5-1}\frac{\delta_{n_5'', r}}{\overline v_0(r+1/2)}\, .\label{Higgs} \end{eqnarray} Then the Higgs mass is extracted according to eq. (\ref{mass}). \section{Spontaneous Symmetry Breaking} In principle one could attempt to massage further the result eq. (\ref{potb}) and in particular consider its small lattice spacing expansion. We will not attempt such a task here, instead we will compute it numerically and fit it to the various possible forms that the static potential could assume. Furthermore, we will present cases where only a four-dimensional Yukawa fit is possible. Having the Standard Model Higgs mechanism in mind, we will compute the quantity \begin{equation} \rho_{HZ} = \frac{m_H}{m_Z}\label{rho} \end{equation} with $m_H$ the mass of the complex scalar and $m_Z$ the mass of the boundary $U(1)$ gauge boson, extracted from the Yukawa fit. This quantity is infinite in the absence of SSB and finite when there is SSB. In our framework it depends on three parameters: $\beta$, $\gamma$ and $N_5$. We fix $\gamma=0.55$ and fix $\beta$ so that $F_1=m_H R$ (with $R$ the length of the interval - fifth dimension) is constant and follow its $N_5$-dependence. In fact, it is possible to extract from the static potential not only the Yukawa mass $m_Z$ but also the mass of the first excited state. Such a state, from the point of view of models beyond the Standard Model is typically called a $Z'$. Thus, we can define a similar to eq. (\ref{rho}) quantity \begin{equation} \rho_{HZ'} = \frac{m_H}{m_{Z'}}\label{rhoprime}. \end{equation} Clearly, by computing both eqs. (\ref{rho}) and (\ref{rhoprime}) and fixing to a definite value one of them, would give us a prediction for the other. The recent result from the LHC motivates us to fix (approximately) \begin{equation} \rho_{HZ} = 1.3875\label{SMrho} \end{equation} which then leaves us with a definite prediction for $m_{Z'}$. The caveat here of course is that such a process would have to be performed along a Line of Constant Physics (LCP) \cite{LCP}. Here we will only present data to show that $\rho_{HZ}$ is finite and therefore argue that there is a non-perturbative dynamical mechanism of SSB in the pure gauge system. We use the term dynamical in order to stress that the gauge boson becomes massive without introducing by hand a vacuum expectation value beyond that of the MF background. On fig. \ref{f_rho_g0p55} we show our main result regarding SSB in the lattice $SU(2)$ orbifold model in the MF expansion, for $\gamma=0.55$ and $F_1=0.2$. Evidently the order parameter that signals SSB is not infinite for a wide range of $N_5$ values, consistent with earlier lattice Monte Carlo results \cite{MC1}. Moreover, preliminary results show that it is possible to get for $\rho_{HZ}$ the value of eq. (\ref{SMrho}), an analysis that will be presented in \cite{LCP}. \begin{figure}\centering \resizebox{10cm}{!}{\includegraphics[angle=0]{rho_L200g0p55F1_0p2_boundary.eps}} \caption{The ratio of the Higgs to the $Z$ and $Z'$ boson masses in the mean-field extracted from the static potential on the boundary.} \label{f_rho_g0p55} \end{figure} \section{Conclusions} We computed the Higgs to $Z$-boson mass ratio in a five-dimensional $SU(2)$ gauge theory regularized on an anisotropic lattice, with the anisotropy pointing along the fifth-dimension. The method is that of an analytical Mean-Field expansion around a non-trivial background, which is evidently a good approximation to the non-perturbative theory in five (or higher) dimensions. We computed this quantity in the vicinity of the bulk phase transition. Contrary to the analogous calculation in the perturbative regime, we find that there is spontaneous symmetry breaking of the boundary $U(1)$ symmetry already in the pure gauge theory. The breaking is dynamical, since no Higgs vacuum expectation value is introduced and is consistent with results from Monte Carlo investigations. {\bf Acknowledgments.} K. Y. is supported by the Marie Curie Initial Training Network STRONGnet. STRONGnet is funded by the European Union under Grant Agreement number 238353 (ITN STRONGnet). N. I. thanks the Alexander von Humboldt Foundation for support. N. I. was partially supported by the NTUA research program PEBE 2010.
\section{Introduction} \label{secintroduction} Stochastic simulation of physical variables such as minimum or maximum temperature, precipitation amount and solar radiation are often required\vadjust{\goodbreak} as inputs to physical models over varying types of topography. Over plains regions, agricultural and crop models require daily minimum and maximum temperature simulations at locations that typically do not have direct observations. In mountainous regions, hydrological models require stochastic weather realizations for runoff, snowmelt and watershed modeling, as well as water resource planning and climate impact assessment [\citet{kustas1994}, \citet{semenov1997}]. Stochastic weather generators (SWGs) are one approach to producing simulations of daily weather; they are simply probability models whose simulations are statistically similar to observations [\citet{wilks1999ppg}]. SWGs can loosely be categorized into model-based [e.g., \citet{racsko1991}, \citet{richardson1981}] and empirical approaches [e.g., \citet{lall1996}, \citet{rajagopalan1999}]. Often these weather generators produce simulations only at locations with observational data, but modern physical models require gridded daily weather. Hence, recent research has been directed toward generating spatially consistent SWGs that are available at and between observation locations [\citet {wilks1999}, \citet{kleiber2012}]. Herein we focus on a model-based approach to minimum and maximum temperature simulation over a mix of complex terrain and relatively homogeneous terrain simultaneously. Spatially consistent simulation over most agricultural regions can be accommodated using isotropic or stationary models that are appropriate for regions with relatively constant or slowly changing topography. Domains with highly variable terrain, in particular, mountainous domains, are challenging for the majority of univariate spatial models due to substantial nonstationarity of physical processes in these areas. Weather over complex terrain is highly variable due to topography; for example, at high elevations in the northern hemisphere, north facing slopes tend to be cooler than lower elevations and south facing slopes and valleys can create their own micro-climate relative to the surrounding high elevation. These conspire to produce intricate spatial variability that is hard for models to capture. A typical approach is to partition the space into homogeneous regions and model each region separately. While a number of statistical nonstationary spatial models have been proposed for univariate fields [\citet{fuentes2002}, \citet{haas1990}, \citet{higdon1998}, \citet{kim2005}, \citet{paciorek2006}, \citet{pintore2006}, \citet{sampson1992}, \citet{stroud2001}], fewer are available for multivariate spatial simulation, which is of key concern for simultaneous minimum and maximum temperature simulation [\citet{gelfand2004}, \citet{jun2011}, \citet{kleiber2012nsmm}, \citet{shaddick2002}]. \begin{figure} \includegraphics{602f01.eps} \caption{Map of elevations in Colorado (in meters) and the 145 locations used from the Global Historical Climatology Network. Four locations we later use for cross-validation are denoted \textup{(a)}~Kit Carson, \textup {(b)} Estes Park, \textup{(c)} Buena Vista and \textup{(d)} Delta.} \label{fignetwork} \end{figure} Some literature in geography and the atmospheric sciences is concerned with deterministic interpolation of observed weather variables, often over domains with complex terrain [\citet{daly1994}, \citet{hijmans2005}, \citet{hutchinson1995}, \citet{legates1990}, \citet{price2000}, \citet{running1987}, \citet{thornton1997}, \citet{willmott1995}]. The common theme among these approaches is the inclusion of high resolution digital elevation maps as well as other physical information such as slope and aspect to deterministically interpolate meteorological variables. While most of these models are sophisticated physical interpolation schemes, they [apart from \citet{thornton1997}] are chiefly concerned with monthly or annual average quantities, and do not produce stochastic realizations of daily weather, which is our primary interest. These schemes are also typically ad hoc, and are not based on a formal statistical model. Stochastic interpolation and simulation of physical variables has persistent interest in the statistics literature. Often, precipitation holds the primary interest, as its mixed discrete-continuous and skewed nature pose substantial challenges [\citet{ailliot2009}, \citet{allcroft2003}, \citet{brown2001}, \citet{durban2001}, \citet{hughes1999}, \citet{sanso2000}]. However, recent authors have acknowledged the difficulties of temperature modeling in complex terrain [\citet{paciorek2006}], and \citet{gelfand2005} is one of few to simultaneously model temperature and precipitation. The study domain in this paper is the state of Colorado. Figure~\ref{fignetwork} illustrates the challenging terrain of Colorado, with eastern plains dipping to a minimum elevation of approximately 1000~m and the Rocky Mountains of central Colorado peaking out at above 4000~m. The front range, the ridge separating the Rocky Mountains from the eastern plains (running north-south on approximately the $-105^{\circ}$ longitude line), is especially difficult to accommodate using the currently available multivariate covariance models, most of which are isotropic models, and do not allow for sudden boundaries or even gradually evolving spatial structures across a domain. The 145 locations shown in Figure~\ref{fignetwork} are a subset of stations from the Global Historical Climatology Network Database [GHCND; \citet{peterson1997}]. Daily observations of minimum and maximum temperatures are available between a time period of at most 1893 through 2011. Associated with each observation is a quality flag provided by the GHCND; we removed all flagged observations to avoid poor quality observations. In this paper we propose a framework for bivariate stochastic temperature simulation that splits the model into two components. The first component represents local climate, allowing the average behavior of minimum and maximum temperature to vary with location, which is of critical concern in regions such as Colorado with the average behavior of temperature in the Rocky Mountains being vastly different than that over the eastern plains. The second component can be interpreted as daily weather, yielding local variability in space and time, and preserving the spatial correlation between both processes. \section{Stochastic model} \label{secmodel} Consider the bivariate process of minimum temperature, $Z_N(\bs,t)$, and maximum temperature, $Z_X(\bs,t)$, at location $\bs\in\real^2$ on day $t=1,\ldots,T$. Our model for the bivariate process is \begin{eqnarray} Z_N(\bs,t) &=& \bbeta_N(\bs)' \bX_N(\bs,t) + W_N(\bs,t), \label{eqNmodel} \\ Z_X(\bs,t) &=& \bbeta_X(\bs)' \bX_X(\bs,t) + W_X(\bs,t) \label{eqXmodel}. \end{eqnarray} The vector of coefficients $\bbeta_i(\bs) = (\beta_{0i}(\bs),\beta _{1i}(\bs), \ldots,\beta_{pi}(\bs))'$, for $i=N,X$, may be of different length for minimum and maximum temperatures, allowing for distinct sets of covariates, although for notational simplicity we assume both processes share the same number of covariates, $p+1$. The covariates $\bX_i(\bs,t) = (X_{0i}(\bs ,t),\ldots,X_{pi}(\bs,t))'$ typically involve autoregressive and seasonality terms and, if available, can contain additional information such as regional climate model output. It is convenient to view the models of (\ref{eqNmodel}) and (\ref{eqXmodel}) as a sum of ``local climate'' plus ``weather.'' The local climate is dependent on spatially and temporally varying covariates, and whose coefficients vary across the domain, allowing for the relative influence of each covariate to depend on location. The weather terms, $W_N(\bs,t)$ and $W_X(\bs,t)$, capture small scale variability and correlate the bivariate temperature process across space. \subsection{Local climate component} \label{secclimate} The coefficients $\beta_{ki}(\bs)$, for $i=N,X$ and $k=0,\ldots,p$, allow the average behavior of temperature to vary with location. This is crucially important in areas of complex terrain or over large domains where variable orography and general circulation patterns give rise to varying climate [\citet{chandler2005}, \citet{johnson2000}, \citet{kleiber2012}]. \citet{pepin2002} point out that climate change trends in Colorado are highly dependent on the terrain. Direct estimates of these coefficients are usually only available at locations within the observation network, so we model the coefficients as spatial Gaussian processes. In particular, we suppose $\beta_{ki}(\bs)$ has mean $\mu_{ki}$ and Mat\'ern covariance augmented with a nugget effect, with variance parameter $\sigma_{ki}^2$, range $a_{ki}$, smoothness $\nu_{ki}$ and nugget effect $\tau_{ki}^2$ [\citet{guttorp2006}]. The goal of a spatial model for the coefficients $\beta_{ki}(\bs)$ is for interpolation from the observational network locations to a chosen grid. The Mat\'ern is an isotropic covariance function that is especially useful for kriging [\citet{stein1999}]. One might consider using a nonstationary function for the coefficient covariance model, but in our experience (see the example section below), the simpler stationary model works well for local climate interpolation. For Colorado, we use the following covariates: \begin{eqnarray} \label{eqcovariates}\qquad \bX_N(\bs,t) = \biggl(1,\cos\biggl( \frac{2\pi t}{365} \biggr), \sin\biggl(\frac{2\pi t}{365} \biggr),Z_X( \bs,t-1), Z_N(\bs,t-1),r_t \biggr)', \end{eqnarray} with the corresponding case for $\bX_X(\bs,t)$ reversing indices $N$ and $X$. The harmonics allow for seasonality in minimum and maximum temperatures, and we include bivariate autoregressive terms to account for temporal persistence of temperature. The final covariate, $r_t$, is a linear drift of length $T$ between $-1$ and 1 (for numerical stability), which we include to control for temperature trends over the 119 year period of our data set, noting that these trends do not necessarily reflect global warming. These covariates were selected using a BIC criterion at all individual stations; that is, fitting a model to each location independently, the model with all of the above covariates had the smallest BIC value for all stations within the GHCND in Colorado, as compared to any subset of the selected covariates. We considered models with higher order harmonics and autoregressive lags, but the results were nearly identical to those presented below, hence, we favor the simpler set of covariates. Suppose we observe the bivariate process $(Z_N(\bs,t),Z_X(\bs,t))'$ at locations $\bs=\bs_1,\ldots,\bs_n$ and time points $t = 1,\ldots,{T}$. At each location within the observation network, we estimate local parameters $\hat{\beta}_{ki}(\bs)$ by ordinary least squares. These estimates have low uncertainty; in the Colorado network, the location with the sparsest observational record still has more than 10,000 available observations. Conditional on the estimates $\hat{\beta}_{ki}(\bs)$, we estimate the spatial Gaussian process parameters $\mu_{ki}, \sigma_{ki}^2, a_{ki}, \nu_{ki}$ and $\tau_{ki}^2$ by maximum likelihood, exploiting the Gaussian process assumption. These spatially varying coefficients models [\citet{gelfand2003}] have been used for probabilistic forecasting, with a similar two-step estimation procedure [\citet{kleiber2011mwr}, \citet{kleiber2011jasa}]. At an arbitrary location $\bs_0$, not necessarily within the observation network, we spatially interpolate the estimates $\hat{\bbeta}_{ki} = (\hat {\beta}_{ki}(\bs_1), \ldots,\hat{\beta}_{ki}(\bs_n))'$ via\vadjust{\goodbreak} kriging [\citet{cressie1993}]. In particular, the kriging estimator is \[ \hat{\beta}_{ki}(\mathbf{s}_0) = \mathbf{c}' \Sigma^{-1} (\hat{\bbeta}_{ki} - \mu_{ki} \mathbf{1}) + \mu_{ki} \] and the interpolation variance is \[ \sigma_{ki}^2 + \tau_{ki}^2 - \mathbf{c}' \Sigma^{-1} \mathbf{c}, \] where $\mathbf{1}$ is a vector of 1s of length $n$, $\mathbf{c}' = (\Cov(\beta_{ki}(\bs _0),\beta_{ki}(\bs_1)),\ldots,\break \Cov(\beta_{ki}(\bs_0),\beta_{ki}(\bs_n)))$ and $(\Sigma)_{j,\ell } = \Cov(\beta_{ki}(\bs_j),\beta_{ki}(\bs_\ell))$ for $j,\ell =1,\ldots,n$. As kriging is an exact interpolator, when $\bs_0 = \bs_\ell$ for any $\ell=1,\ldots,n$, the interpolator returns the ordinary least squares estimate $\hat{\beta}_{ki}(\bs_\ell)$. In the next section we exploit a nonparametric estimator of the covariance function for the bivariate weather process. Key to the nonparametric estimator being consistent is a large number of realizations of the process [\citet{kleiber2012nsmm}] which we have available for the residual weather processes, whereas the coefficient processes of the local climate component have only one realization. Hence, we favor the parametric model with a two-step estimation procedure for local climate. \subsection{Weather component} \label{secweather} To simulate spatially correlated fields of minimum and maximum temperatures consistent with observed spatial patterns, we require a bivariate spatial model for $W_N(\bs,t)$ and $W_X(\bs,t)$. In particular, we model these weather processes as a zero-mean bivariate spatial Gaussian process indexed by day of the year. For locations $\mathbf{x},\by$, and arbitrary time point $t$, the bivariate covariance model is \begin{eqnarray} &\Cov\bigl(W_i(\mathbf{x},t),W_j(\by,t+1)\bigr) = 0, \label{eqWtime} \\ &\Cov\bigl(W_i(\mathbf{x},t),W_i(\by,t)\bigr) = C_{ii}\bigl(\mathbf{x},\by,d(t)\bigr) + \tau _i(\mathbf{x}, \by)^2 \indicator_{[\mathbf{x} = \mathbf{y}]}, \label{eqWcov} \\ &\Cov\bigl(W_i(\mathbf{x},t),W_j(\by,t)\bigr) = C_{ij}\bigl(\mathbf{x},\by,d(t)\bigr)\qquad \mbox{for } i\not=j\label{eqWcross} \end{eqnarray} for $i,j=N,X$, where $d(t) \in\{1,\ldots,365\}$ is just the calendar day of the year on which time point $t$ falls. The covariance model of (\ref{eqWtime}), (\ref{eqWcov}) and (\ref{eqWcross}) implies some important assumptions. First, we assume temporal dependence has been accounted for in the local mean function (e.g., via autoregressive terms) so that the weather process is temporally independent, hence (\ref{eqWtime}). Indeed, exploratory plots such as autocorrelation functions and empirical covariance functions indicate the bivariate autoregression of (\ref{eqcovariates}) is sufficient to account for the temporal persistence of temperature in Colorado; see the example section below. Second, the covariance and cross-covariance functions $C_{ii}(\mathbf{x},\by,d(t))$ and $C_{ij}(\mathbf{x},\by,d(t))$ depend on the day of year, allowing the bivariate process to have seasonally dependent second-order structure. In (\ref{eqWcov}), $\tau_i(\mathbf{x})^2 = \tau _i(\mathbf{x},\mathbf{x})^2$ is a local nugget effect, accounting for small scale variability as well as measurement error. In the geostatistical literature, $C_{ii}(\mathbf{x},\mathbf{x},d(t))$ is often termed\vadjust{\goodbreak} the marginal variance, while $C_{ii}(\mathbf{x},\mathbf{x},d(t)) + \tau_i(\mathbf{x})^2$ is called the sill, that is, the total variance at a given location [\citet{cressie1993}]. Unlike most geostatistical models [\citet{christensen2011} being a notable departure], we allow the nugget effect to vary with location, as we expect the small scale variability to be highly dependent on orography. At any fixed time point $t$ [i.e., calendar day $d(t)$], we require the matrix-valued covariance function \begin{equation} \bC\bigl(\mathbf{x},\by,d(t)\bigr) = \pmatrix{ C_{NN} \bigl(\mathbf{x},\by,d(t)\bigr) & C_{NX}\bigl (\mathbf{x},\by,d(t)\bigr) \vspace*{2pt}\cr C_{XN}\bigl(\mathbf{x},\by,d(t)\bigr) & C_{XX}\bigl(\mathbf{x},\by,d(t) \bigr)} \label{eqmatrixcov} \end{equation} to be a nonnegative definite matrix function. Specifically, at arbitrary locations $\bs_1,\ldots,\bs_n$, the covariance matrix of the random vector \[ \bigl(W_N(\bs_1,t),W_X( \bs_1,t),W_N(\bs_2,t),W_X( \bs_2,t),\ldots,W_N(\bs_n,t),W_X( \bs_n,t)\bigr)', \] which is made up of blocks $\bC(\bs_k,\bs_\ell,d(t))$, must be nonnegative definite. Over regions with complex terrain, temperature observations can exhibit substantial nonstationarity [\citet{paciorek2006}]. While some multivariate spatial models that can account for nonstationarity are available [e.g., \citet{gelfand2004}, \citet{kleiber2012nsmm}], these are parametric models with locally varying parameter functions that are difficult to estimate. We aim to exploit the large number of replications and reasonably well covered observation network of the GHCND over Colorado, and propose a nonparametric estimator of the matrix-valued covariance function that retains nonnegative definiteness. In particular, suppose the bivariate process is observed at locations $\bs_k, k=1,\ldots,n$, and times $t=1,\ldots,T$. Then our nonparametric estimator of $C_{ij}(\mathbf{x},\by,d(t_0))$ in (\ref{eqWcov}) and (\ref{eqWcross}), at location pair $(\mathbf{x},\by)$ and time point $t_0$ is \begin{eqnarray} \label{eqnp} &&\hat{C}_{ij}\bigl(\mathbf{x},\by ,d(t_0)\bigr) \nonumber\\ &&\qquad =\Biggl(\sum_{t=1}^{T} \sum_{k=1}^n \sum_{\ell=1}^n K_{\lambda_t} \bigl(\bigl\|d(t_0),d(t)\bigr\|_d \bigr) K_\lambda\bigl(\|\mathbf{x}-\bs_k\| \bigr) K_\lambda\bigl(\|\by-\bs_\ell\| \bigr) \nonumber \\[-8pt] \\[-8pt] \nonumber &&\hspace*{169pt}\qquad\quad{}\times W_i(\bs_k,t) W_j(\bs_\ell,t)\Biggr)\\ &&\qquad\quad{}\bigg/\Biggl(\sum_{t=1}^{T} \sum_{k=1}^n \sum _{\ell=1}^n K_{\lambda_t}\bigl(\bigl\|d(t_0),d(t)\bigr\|_d\bigr) K_\lambda\bigl(\|\mathbf{x}-\bs_k\|\bigr) K_\lambda\bigl(\|\by-\bs_\ell\|\bigr)\Biggr)\nonumber \end{eqnarray} for $i,j=N,X$. Here, $K_\lambda$ is a kernel function with bandwidth $\lambda$, and we use $K_\lambda(\|\bh\|) = (1/\lambda)\exp(-\|\bh\|/\lambda)$. We use the Euclidean norm $\|\cdot\|$, and the distance function $\|\cdot,\cdot\|_d$ is the distance between days of the year so that $\| d_1,d_2\|_d = |d_1-d_2|$ for $|d_1-d_2|\leq182$ and $\| d_1,d_2\|_d = |365-|d_1-d_2||$ for $|d_1-d_2|>182$, where $d_1,d_2=1,\ldots,365$,\vadjust{\goodbreak} for example, $\| 1, 365\|_d = 1$. Occasionally $Z_i(\bs_k,t)$ [and subsequently $W_i(\bs_k,t)$] is not available in practice due to instrument failure or disruptions in communications. The estimator we use operationally is a slightly modified version of (\ref{eqnp}), where we make the convention $W_i(\bs,t) \indicator_{[W_i(\mathbf{s},t)\ \mathrm{is\ observed}]} = 0$ when $W_i(\bs,t)$ is missing. It is convenient to define the single-time-point smoothed empirical covariance function\looseness=-1 \begin{eqnarray} \label{eqnpR} &&\hat{R}_{ij}(\mathbf{x},\by,t) \nonumber\\[-3pt] &&\qquad= \Biggl(\sum_{k=1}^n \sum_{\ell=1}^n K_\lambda\bigl(\|\mathbf{x}-\bs_k\| \bigr)K_\lambda\bigl(\|\by -\bs_\ell\| \bigr) W_i(\bs_k,t) W_j(\bs_\ell,t)\nonumber\\[-3pt] &&\hspace*{94pt}{}\times \indicator_{[W_i(\mathbf{s}_k,t)\ \mathrm{is\ observed}]} \indicator_{[W_j(\mathbf{s}_\ell,t)\ \mathrm{is\ observed}]}\Biggr)\\[-3pt] &&\qquad\quad{}\bigg/ \Biggl( \sum_{k=1}^n \sum_{\ell=1}^n K_\lambda\bigl(\|\mathbf{x}-\bs_k\|\bigr) K_\lambda\bigl(\|\by -\bs_\ell\|\bigr)\nonumber\\[-3pt] &&\hspace*{50pt}\qquad\quad{}\times \indicator_{[W_i(\mathbf{s}_k,t)\ \mathrm{is\ observed}]} \indicator_{[W_j(\mathbf{s}_\ell,t)\ \mathrm{is\ observed}]}\Biggr).\nonumber \end{eqnarray}\looseness=0 Notice $\hat{R}_{ij}(\mathbf{x},\by,t)$ is just a (spatially) smoothed empirical covariance function over the available observations on day $t$. $\hat{R}_{ij}(\mathbf{x},\by,t)$ is a nonnegative definite multivariate covariance function, a property we show in the \hyperref[app]{Appendix}. Our adjusted version of (\ref{eqnp}) that accounts for missing observations then is \begin{equation} \label{eqnpmissing} \hat{C}_{ij}\bigl(\mathbf{x},\by,d(t_0) \bigr) = \frac{\sum_{t=1}^T K_{\lambda_t} (\|d(t_0),d(t)\|_d ) \hat{R}_{ij}(\mathbf{x},\by,t)} { \sum_{t=1}^{T} K_{\lambda_t}(\|d(t_0),d(t)\|_d)}. \end{equation} As $\hat{C}_{ij}(\mathbf{x},\by,d(t_0))$ is a positively weighted linear combination of multivariate covariance functions, it is again nonnegative definite. The estimator for missing observations (\ref{eqnpmissing}) reduces to the original estimator (\ref{eqnp}) when no observations are missing and, hence, (\ref{eqnp}) is also nonnegative definite. The estimator (\ref{eqnp}) is a smoothed version of daily empirical covariance matrices. The first level of smoothing yields an estimate of spatial covariance at any arbitrary location pairs in the domain. The temporal smoothing shares information between adjacent time points, where we assume that spatial covariance on a given day is similar to that in a short period leading up to that day, and in a short period following that day. This estimator is a generalization of kernel smoothed empirical covariance estimators considered by \citet{oehlert1993}, \citet{guillot2001} and \citet{jun2011mwr} to the multivariate process setting evolving across time. We estimate the time bandwidth $\lambda_t$ by predictive leave-one-out cross-validation, leaving out local empirical variance estimates. The estimated bandwidth for time is $\hat{\lambda}_t = 7.8$ days. We use cross-validation for the temporal bandwidth, as we assume the temporal evolution of spatial covariance is\vadjust{\goodbreak} slowly evolving across time, for example,~we do not expect a sharp change in spatial covariance between June~1 and June~2. In our experience, using cross-validation for the spatial bandwidth parameter $\lambda$ oversmooths the spatial covariance function. When kernel smoothing a mean function, cross-validation is generally acknowledged to yield more variability than is expected for a smoothly varying mean function, and typically the bandwidth must be inflated [\citet{wand1995}]. However, this experience is under the assumption that the mean function is varying smoothly across the domain, and in regions of complex terrain we expect the opposite behavior, where sharp boundaries of the covariance function may exist due to sudden changes in elevation. For example, cross-validation implies the optimal spatial bandwidth is 75~km, which implies an effective range of the kernel function (i.e., up to $5\%$ weight) of approximately 225 km, greatly oversmoothing regions such as the San Luis Valley in southern Colorado, at approximately 100 km across. Hence, we choose a bandwidth such that the effective distance of the kernel function coincides with the $5\%$ quantile of all intersite distances (62 km); the heuristic argument is that, for approximately evenly distributed observation locations, the covariance estimator at a given location uses the nearest $5\%$ of available network locations and down-weights remote locations; this ad hoc criterion implies a spatial bandwidth of $\hat{\lambda} = 22$~km. The estimator (\ref{eqnp}) is asymptotically unbiased for $C_{ij}(\mathbf{x},\by,d(t_0))$ when the domain sample size increases and the bandwidth decreases to zero sufficiently quickly. A short argument is given in the \hyperref[app]{Appendix}. In fact, it can be shown that the estimator is consistent for $C_{ij}(\mathbf{x},\by,d(t_0))$, using arguments similar to those of \citet{kleiber2012nsmm}, but this is beyond the scope of the present paper. All that remains to be estimated is the local nugget effect $\tau _i(\bs)^2$. At each observation location $\bs_k, k=1,\ldots,n$, and time point $t=1,\ldots,T$, let $W_i(\bs_k,t)$ be the estimated residual $Z_i(\bs _k,t) - \hat{\bbeta}_i(\bs_k)'\bX_i(\bs_k,t)$. Define the local empirical variance on day $d=1,\ldots,365$ as \[ \hat{\sigma}_{i}(\bs_k,d)^2 = \frac{1}{\# \{ t | d(t) = d \}} \sum_{\{ t | d(t) = d \}} W_i( \bs_k,t)^2, \] where $\#$ denotes cardinality of the set, with the natural redefinition for missing values of $W_i(\bs_k,t)$. Intuitively, a good estimator for $\tau_i(\bs_k)^2$ is \begin{equation} \label{eqtau} \hat{\tau}_i(\bs_k)^2 = \frac{1}{365} \sum_{d=1}^{365} \bigl(\hat{ \sigma}_i(\bs_k,d)^2 - \hat{C}_{ii}( \bs_k,\bs_k,d) \bigr), \end{equation} since, by the law of large numbers, $\hat{\sigma}_{i}(\bs_k,d)^2 \rightarrow C_{ii}(\bs_k,\bs_k,d) + \tau_i(\bs_k)^2$, where the convergence is taken as $T\rightarrow\infty$, and by the argument in the \hyperref[app]{Appendix}, $\hat{C}_{ii}(\bs_k,\bs_k,d) \rightarrow C_{ii}(\bs_k,\bs_k,d)$. While theoretically appealing, in practice, due to the smoothing in $\hat{C}_{ii}$, at some locations the estimate $\hat{\tau}_i(\bs _k)$ is negative. Hence, in similar spirit we use (\ref{eqtau}), but set the invalid estimates to zero.\vadjust{\goodbreak} Estimates of $\tau_{i}(\bs)^2$ are gathered at arbitrary locations, that is, not necessarily within the observation network, by imposing a probabilistic spatial structure on $\tau_i(\bs)$. In particular, we model $\tau_i(\bs)$ as a Gaussian process with spatially constant mean and Mat\'ern covariance function, augmented with a nugget effect. Just as for the spatial parameters of the $\beta_{ki}(\bs)$, we estimate the spatial parameters of $\tau_i(\bs)$ by maximum likelihood, conditional on the estimates $\{\hat{\tau}_i(\bs_k)\}_{k=1}^n$. While the estimates $\hat{\tau}_i(\bs_k)$ at observation locations are always valid, the kriging interpolator of $\tau_i(\bs)$ may occasionally take on very small negative values; in our example below we did not experience such an issue, but in other domains these degenerate estimates may be artificially set to zero. \section{Minimum and maximum temperature in Colorado} \label{secexample} We fit our model to the data from the 145 GHCND locations shown in Figure~\ref{fignetwork}. For simplicity, we removed all leap days from the 119 years of available data, so that each year has 365 days. Using all available data, we fit local climate parameters by ordinary least squares and estimate temporally varying multivariate spatial covariances using the nonparametric estimator (\ref{eqnpmissing}) applied to the observed residuals. We then simulate the bivariate process for a 119 year trajectory to compare to the observed bivariate series. The first day's (January 1, 1893) simulation requires autoregressive terms in~(\ref{eqNmodel}) and~(\ref{eqXmodel}); we initialize using the climatological domain average of minimum and maximum temperatures on December~31. The resulting simulations are masked to share the same missing value pattern as the observations. \begin{figure} \includegraphics{602f02.eps} \caption{Empirical autocorrelation functions for minimum and maximum temperature residuals at locations \textup{(a)} Kit Carson, \textup{(b)} Estes Park, \textup{(c)} Buena Vista and \textup{(d)} Delta.}\label{figacfs} \end{figure} Recall the assumption implied by equation (\ref{eqWtime}), where we assume temporal dependence has been accounted for in the local mean function via the bivariate autoregression. Figure~\ref{figacfs} contains empirical autocorrelation functions for the observed residuals $W_N(\bs,t)$ and $W_X(\bs,t)$ at four network stations, shown in Figure~\ref{fignetwork}. These locations we view as representative of four distinct regimes of Colorado: eastern plains (a, Kit Carson), front range (b, Estes Park), Rocky Mountains (c, Buena Vista) and the western slopes (d, Delta). It is evident that the bivariate autoregression accounts for the majority of temporal persistence in temperature; the maximal lag-1 autocorrelation coefficient for the residual processes at these four stations is $0.06$ at Estes Park, whereas all other coefficients are less than or equal to $0.03$. \begin{figure} \includegraphics{602f03.eps} \caption{Scatterplots comparing empirical pairwise station correlation to simulated correlations using a bivariate stationary model (grey dots) or the nonstationary nonparametric model (black dots) over the summer months (JJA). The diagonal line indicates perfect agreement between model and empirical correlations.}\label{figpairwise} \end{figure} To motivate the flexibility of the nonparametric estimator (\ref{eqnp}), we compare it to a state-of-the-art isotropic bivariate spatial model. In particular, we fit a bivariate Mat\'ern model [\citet{gneiting2010}, \citet{apanasovich2012}] augmented with a nugget effect, where \begin{eqnarray}\quad C_{ii}(\mathbf{x},\by,t) &=& \frac{1}{2^{\nu _i-1}\Gamma(\nu_i)} \bigl(a_i\|\mathbf{x}-\by \|\bigr)^{\nu_i} \mathrm{ K}_{\nu_i}\bigl(a_i\|\mathbf{x}-\by\| \bigr) + \tau_i^2\indicator_{[\mathbf{x}=\mathbf{y}]} , \\ \quad\qquad C_{NX}(\mathbf{x},\by,t) &= &\rho_{NX} \frac {1}{2^{\nu_{NX}-1}\Gamma(\nu_{NX})} \bigl(a_{NX}\|\mathbf{x}-\by\|\bigr)^{\nu_i} \mathrm{ K}_{\nu _{NX}}\bigl(a_{NX} \|\mathbf{x}-\by\|\bigr)\vadjust{\goodbreak} \end{eqnarray} for $i=N,X$, where $\mathrm{ K}_\nu$ is a modified Bessel function of the second kind of order $\nu$, and $\nu_{NX} = (\nu_N+\nu_X)/2$ and $a_{NX} = \min(a_N,a_X)$. We fit the parameters by maximum likelihood, viewing each bivariate estimated residual $(W_N(\bs,t),W_X(\bs,t))$ as independent across time. In the stochastic weather simulation literature, it is customary to fit separate models for each season. While our nonparametric estimator is available on any day, to facilitate comparisons to the bivariate Mat\'ern, we fit both models to only the summer months (JJA), and compare empirical to simulated correlations and cross-correlations under both the isotropic and nonparametric models; Figure~\ref{figpairwise} displays these results. The stationary model tends to overestimate spatial correlation for both minimum and maximum temperatures, whereas our nonstationary model adequately captures low and high correlations simultaneously. The third panel of Figure~\ref{figpairwise} shows empirical against simulated cross-correlations. Substantial nonstationarity of cross-correlation across Colorado is well modeled by our nonparametric approach, but the stationary model clearly fails, putting most cross-correlations at around $0.10$, whereas the empirical estimates suggest the true cross-correlations should vary between $-0.10$ and $0.40$. \begin{figure} \includegraphics{602f04.eps} \caption{Plots of spatial correlation and cross-correlation functions, $C_{ij}(\bs_0,\cdot,d(t))$, on $d(t)=$ June~1 where $\bs_0$ is a grid location in the eastern plains (top row) or a grid location in the Rocky Mountains (bottom row), with grid locations indicated by black dots. Each pixel's color indicates the model estimated spatial correlation between the pixel location and the dot.} \label{figspatialcor}\vspace*{-3pt} \end{figure} Our nonparametric matrix covariance estimator (\ref{eqnp}) accommodates nonstationary behavior of the multivariate process. Figure~\ref{figspatialcor} shows two covariance functions on June 1, one whose first argument is based at a grid location in the eastern plains of Colorado, and the second covariance function whose first argument is based at a grid location in the Rocky Mountains. The top row is the covariance function for the plains-based grid location; particularly for maximum temperature, and lesser so for minimum temperature, there is strong positive within variable correlation throughout the plains region, suggesting that maximum temperatures are highly correlated across the plains. At the front range boundary (approximately $-105^\circ$ longitude), there is a sharp drop off in spatial correlation from approximately $0.80$ over the plains to $0.40$ in the Rocky Mountains. This is due to the fact that temperature is more highly correlated within the two main types of topography of Colorado, either the plains or mountains, but not between the two types. Hence, our estimator is able to capture the sharp boundary between the eastern plains and Rocky Mountains for within variable spatial correlation. Our estimator also identifies the positive cross-correlation between minimum and maximum temperatures in the plains, but allows the two processes to be effectively independent over the Rocky Mountains.\vadjust{\goodbreak} This nonstationarity of cross-correlation is very difficult to accommodate using extant models, and has only been recently acknowledged in the literature [\citet{kleiber2012nsccc}]. \begin{figure}[b]\vspace*{-3pt} \includegraphics{602f05.eps} \caption{Same as Figure \protect\ref{figspatialcor}, except for $d(t)=$ January 1 instead of June 1.} \label{figspatialcorjan} \end{figure} Not only does our estimator allow for substantial nonstationarity, the amount and type of nonstationarity is allowed to vary across time. Figure~\ref{figspatialcorjan} shows the same plots\vadjust{\goodbreak} of spatial direct and cross-correlation on January 1, during winter, as opposed to the summer estimates of Figure~\ref{figspatialcor}. In terms of direct covariance, we see the length scale of minimum temperature correlation drastically increase for both the plains- and mountain-based grid locations. In the plains, the spatial correlation structure of maximum temperature is similar during both the winter and summer; on the other hand, this spatial correlation in the mountainous region over winter has a substantially different pattern than over summer. The correlation structure of the weather component for maximum temperature in the Rocky Mountains is clearly nonstationary, implying lower correlation between the example grid point and the southwestern slopes of the Rockies, but having higher correlation along a northwest to southeast transect along the western slopes and through the Rocky Mountains; this pattern makes sense climatologically, as the band of high correlation connects the low lying western Grand Valley area through the lower mountains north of the San Juan chain to the San Luis Valley in southern Colorado. A~similar pattern is present for the cross-correlation function, which is distinct from the summer behavior which indicated near-independence between minimum and maximum temperatures over the complex topography. \begin{figure} \includegraphics{602f06.eps} \caption{Estimated nuggets $\hat{\tau}_N(\bs)$ and $\hat{\tau }_X(\bs)$ at observation network locations, units are degrees Celsius.} \label {fignuggets} \end{figure} A notable departure of our model from typical geostatistical approaches is in allowing the nugget effect to vary with location. Our motivation is that the small scale spatial structure is expected to be dampened in the eastern plains with stable orography, but potentially inflated over the mountainous region of Colorado. Figure~\ref{fignuggets} displays the local estimates $\hat{\tau}_i(\bs)$ for $i=N,X$ at the locations within our observation network. For both minimum and maximum temperatures, the nugget effects tend to be less over the eastern plains, indicating less fine scale spatial structure (although there is yet some evidence of small scale structure in the maximum temperature nuggets here). Over the Rocky Mountains, especially the northern Rockies, minimum temperature exhibits inflated nugget effects, indicating fine scale spatial processes in the complex terrain. Similarly, the finest scale spatial structures indicated by these nugget effects for maximum temperature fall almost directly along the front range, the longitude line of approximately $-105^\circ$, indicating highly variable maximum temperatures between the boundary of the plains and sudden mountainous terrain. The inclusion of a spatially varying $\tau_i(\bs)$ allows the statistical model to retain increased variability along the front range, for example, while simultaneously generating tempered fields over the eastern plains and fields of medium variability over the main Rockies and western slopes. \begin{figure} \includegraphics{602f07.eps} \caption{Q--Q plots for daily spatial extrema, comparing \textup{(a)} domain minimum of minimum temperature, \textup{(b)} domain maximum of maximum temperature, \textup{(c)} domain maximum of minimum temperature and \textup{(d)} domain minimum of maximum temperature, units are degrees Celsius.} \label{figextrema}\vspace*{6pt} \end{figure} An increasingly important consideration in climate science is the effect of climate change on extremes [\citet{easterling2000}]. Our model is not explicitly designed to replicate extreme events, as we focus mainly on the first and second order properties of minimum and maximum temperatures. Figure~\ref{figextrema} shows Q--Q plots for daily domain-wide extrema. In particular, we find the minimal and maximal domain-wide temperatures $Z_{i,\min}(t)=\min_{\mathbf{s}}\{Z_i(\bs,t)\}$ and $Z_{i,\max}(t)=\max_{\mathbf{s}}\{Z_i(\bs,t)\}$, and compare simulated to observed daily statistics for $i=N,X$. Our model replicates the statistical properties of $Z_{N,\min}(t), Z_{X,\min}(t)$ and $Z_{N,\max}(t)$ very well, at even the most extreme tails of these domain extrema. However, we simulate domain-wide maximal maximum temperatures that are slightly too high, on average about $2^\circ$C. Overall, even though our approach does not explicitly model extreme temperatures, we are able to capture the spatial extrema with reasonable accuracy. \begin{figure} \includegraphics{602f08.eps} \caption{Log frequency of observed and simulated local residual threshold exceedances. Each bar's height is the log freqeuency (i.e.,~log number of days) that an exact number of the observation network stations had weather that exceeded the local $90\%$ quantile for \textup{(a)} maximum temperature, or whose weather fell below the local $10\%$ quantile for \textup{(b)} minimum temperature.}\label{figexceedance} \end{figure} While our model adequately replicates domain-wide extrema, the related quantity of spatially consistent local extrema is critically important to replicate. In particular, for energy use forecasting and modeling, if a large number of locations experience unusually low or high temperatures simultaneously, then the load on the energy grid can be much greater than if the temperature anomaly were highly localized. Figure~\ref{figexceedance} shows log frequencies (i.e.,~the log number of days) of the number of stations whose local weather process $W_i(\bs,t)$ either exceeded the local $90\%$ quantile (i.e., the quantile using only data from location $\bs$) or fell below the local $10\%$ quantile, corresponding to local hot or cold events, respectively. Our approach captures the spatial frequencies of unusual local cold temperatures extremely well, and tends to simulate local heat events over slightly inflated regions when many stations experience hot events, although usually fewer than seven extra days on average. \begin{figure} \includegraphics{602f09.eps} \caption{Locally estimated standard deviations $(\hat{C}_{ii}(\bs,\bs,d)^{1/2})$ for $i=N,X$ on all days of the calendar year $d=1,\ldots,365$, and predicted standard deviations for the four hold out stations $\bs= $ \textup{(a)} Kit Carson, \textup{(b)} Estes Park, \textup{(c)} Buena Vista and \textup{(d)} Delta.} \label{figSDCV} \end{figure} Our nonparametric weather component covariance estimator (\ref{eqnp}) is not optimized for cross-validation. To assess the interpolative properties of our estimator, we hold out data from the four network stations shown in Figure~\ref{fignetwork}, representing four distinct regimes of Colorado. We predict the local standard deviations $\hat{C}_{ii}(\bs,\bs,d)^{1/2}$ for $i=N,X$ and compare these to the locally estimated values of $\hat{C}_{ii}(\bs,\bs,d)^{1/2}$ when station data is retained. Figure~\ref{figSDCV} contains the local and predicted estimates for all days of the calendar year. Clearly the weather component variability is highly dependent on season as well as location, particularly for maximum temperature there is substantially greater variability in the eastern plains ($3^\circ$--$5^\circ$C) compared to the mountain regions ($2^\circ$--$3.5^\circ$C). Our predictive local standard deviations (dashed lines in Figure~\ref{figSDCV}) generally agree closely with the local estimates, although there is a slight tendency to under-predict local standard deviation at Kit Carson by $0.1^\circ$--$0.3^\circ$C. Not only are the raw values well predicted, but the climatological curvature is preserved as well; for example, we successfully replicate the increased variability of maximum temperature over the western slopes during springtime with relatively constant variability throughout the three remaining seasons (panel d) while simultaneously producing significant seasonality over the eastern plains, with low variability during summer and high variability during winter (panel a). \begin{table} \tabcolsep=0pt \caption{Interpolated estimates (with predictive standard deviation) of the local climate component coefficients with the validating locally estimated parameters. Locations are $\bs= $ \textup{(a)}~Kit Carson, \textup{(b)}~Estes Park, \textup{(c)}~Buena Vista and \textup{(d)}~Delta. Predictions are starred if the truth is outside of the predictive $95\% $ confidence interval. Units are degrees Celsius for $\beta_0, \beta_1$ and $\beta _2$, unitless for $\beta_3$ and $\beta_4$, and degrees Celsius per century for $\beta _5$}\label{tabests} \begin{tabular*}{\textwidth}{@{\extracolsep{\fill }}ld{2.9}d{2.8}d{2.9}d{2.8}d{2.2}d{2.2}d{2.2}d{2.2}@{}} \hline & \multicolumn{4}{c}{\textbf{Kriged estimate (kriging standard deviation)}} & \multicolumn{4}{c@{}}{\textbf{Local estimate}} \\[-6pt] & \multicolumn{4}{c}{\hrulefill} & \multicolumn{4}{c@{}}{\hrulefill } \\ & \multicolumn{1}{c}{\textbf{a}} & \multicolumn{1}{c}{\textbf{b}} & \multicolumn{1}{c}{\textbf{c}} & \multicolumn{1}{c}{\textbf{d}} & \multicolumn{1}{c}{\textbf{a}} & \multicolumn{1}{c}{\textbf{b}} & \multicolumn{1}{c}{\textbf{c}} & \multicolumn{1}{c@{}}{\textbf{d}} \\ \hline $\beta_{0N}(\bs)$ & -2.70\ (1.40) & -5.43\ (1.32) & -6.37\ (1.35) & -4.39\ (1.34) & -3.43 & -4.76 & -4.58 & -4.49 \\ $\beta_{1N}(\bs)$ & -3.94\ (0.45) & -2.45\ (0.41) & -2.89\ (0.43) & -1.78\ (0.42) & -4.48 & -2.01 & -2.66 & -1.69 \\ $\beta_{2N}(\bs)$ & -1.08\ (0.25) & -0.63\ (0.23) & -0.81\ (0.24) & -0.36\ (0.24) & -1.07 & -0.68 & -0.58 & -0.19 \\ $\beta_{3N}(\bs)$ & 0.20\ (0.06) & 0.28\ (0.06) & 0.27\ (0.06) & 0.27\ (0.06) & 0.20 & 0.27 & 0.21 & 0.25 \\ $\beta_{4N}(\bs)$ & 0.46\ (0.06) & 0.41\ (0.06) & 0.47\ (0.06) & 0.50\ (0.06) & 0.45 & 0.38 & 0.49 & 0.52 \\ $\beta_{5N}(\bs)$ & 0.55\ (1.30) & 0.53\ (1.30) & 0.57\ (1.30) & 0.60\ (1.30) & 0.42 & 0.67 & 0.58 & 0.72 \\[3pt] $\beta_{0X}(\bs)$ & 6.99\ (1.08) & 4.40\ (0.98) & 3.27\ (1.02) & 4.31\ (1.01) & 7.28 & 4.03 & 3.62 & 4.36 \\ $\beta_{1X}(\bs)$ & -4.30^*\ (0.41) & -3.80\ (0.38) & -3.81^*\ (0.39) & -3.77\ (0.39) & -5.29 & -3.30 & -3.00 & -4.08 \\ $\beta_{2X}(\bs)$ & -1.26^*\ (0.15) & -1.31\ (0.14) & -1.29^*\ (0.14) & -0.93\ (0.14) & -1.58 & -1.20 & -0.93 & -0.90 \\ $\beta_{3X}(\bs)$ & 0.03\ (0.07) & -0.01\ (0.07) & -0.02\ (0.07) & -0.03\ (0.07) & -0.05 & -0.01 & -0.02 & -0.08 \\ $\beta_{4X}(\bs)$ & 0.63\ (0.05) & 0.67\ (0.05) & 0.69\ (0.05) & 0.75\ (0.05) & 0.64 & 0.70 & 0.75 & 0.78 \\ $\beta_{5X}(\bs)$ & 0.30\ (0.80) & 0.31\ (0.80) & 0.31\ (0.80) & 0.30\ (0.80) & -0.42 & 0.67 & 0.48 & 0.13 \\ \hline \end{tabular*} \end{table} Table~\ref{tabests} shows the interpolated coefficients with predictive standard deviation, along with the locally estimated parameters $\hat {\beta}_{ki}(\bs)$ for $k=0,\ldots,5$ and $i=N,X$ for $\bs$ being one of the four held out network stations. All locally estimated parameters are within the $95\%$ predictive confidence interval, except for four cases for maximum temperature. Our predictive intervals are calibrated; the coverage of the $95\%$ interpolation intervals for leave-one-station-out cross-validation over all locations was, at worst, $92.4\%$ for $\beta_{2X}(\bs)$. Notice that the local estimates vary substantially between locations, indicating that indeed the local climate varies over the domain. Hence, we are able to successfully predict the local weather component parameters and local climate component parameters at these four hold out locations which are representative of four regimes in Colorado. \begin{figure} \includegraphics{602f10.eps} \caption{Gridded simulation of daily minimum and maximum temperatures on days June~1--4.} \label{figsim} \end{figure} Finally, we illustrate the final product of our approach in Figure \ref{figsim} which displays four days of gridded simulations of minimum and maximum temperatures over Colorado. Marginally, we visually see the temporal persistence of temperature over a period of days, as both minimum and maximum temperatures experience a period of cooling over June 1--4. Notice the effect of local climate is to keep the Rocky Mountain region cooler for both variables, allowing higher minimum and maximum temperatures to fall over the eastern plains of Colorado. We also see slightly warmer temperatures on the western slopes, as the Rocky Mountains decay in elevation to the western border of Colorado. The cross-correlation between the two variables is also present, as both variables are seen to cool across the domain simultaneously. \section{Discussion} \label{secdiscussion} In this paper we introduce a framework for stochastic bivariate minimum and maximum temperature simulation over complex domains. The framework distinguishes between local climate and weather processes. The local climate is accommodated through a linear model whose coefficients are spatially varying, and the weather process is modeled as a bivariate spatial Gaussian process with a nonparametric estimate of the matrix-valued covariance function that retains nonnegative definiteness at arbitrary locations. We successfully capture the temporally varying spatial dependence between minimum and maximum temperatures over the state of Colorado, which exhibits challenging complex terrain that is difficult for extant models to accommodate. Our nonparametric estimator smooths multivariate spatial covariance over space as well as time. This approach allows spatial dependence to be highly different during winter than during summer, for instance, and also retains nonstationary spatial structures both within each process and between processes. The estimator is available at any location, not only those within the observation network, and always retains nonnegative definiteness, allowing for gridded simulations. The estimator relies on kernel-smoothed empirical covariance functions, and our current approach to spatial bandwidth selection is ad hoc. One future route of research may be to decide on a quantitative approach to bandwidth selection when sharp boundaries and highly variable covariances are expected across the study domain, notably different than most mean function smoothing literature [\citet{wand1995}]. A second potential direction of research may be to develop a nonparametric kernel-smoothed estimate of the multivariate covariance function that is robust against outliers and still retains nonnegative definiteness. While our approach does not explicitly model extremes, our simulations indicate reasonable replication of tail behavior, even domain-wide extrema. A limitation in using Gaussian processes is that there is a lack of clustering at high levels, both spatially and temporally [\citet{sibuya1960}]. This is one potential explanation for the behavior of Figure~\ref{figextrema}(b), where domain-wide maximal maximum temperatures were simulated slightly above the observed extremes, although we would expect to see similar behavior in panels (a), (c) and (d). An approach that includes a Gaussian process model for the bulk of the distribution along with a model for spatial extremes may improve extremal performance. One consideration of our model is that we do not explicitly force the simulation of maximum temperature to be greater than or equal to minimum temperature; in our Colorado example maximum temperature was less than minimum temperature for approximately one tenth of a percent of our simulations. It may be of interest to adopt the models of \citet{jolliffe1996} or \citet{jones2004} to our situation if this issue is of critical concern. The clearest route of future research is to extend our ideas to a full stochastic weather simulator that can simulate spatially correlated fields of multiple variables such as minimum and maximum temperatures, precipitation amount, solar radiation, wind direction/wind speed and relative humidity simultaneously. Indeed, in complex terrains the practitioner will need to rely on highly flexible spatial models to replicate the strong nonstationarities exhibited by these various processes, as well as the complicated spatially evolving relationship between them. \begin{appendix} \section*{Appendix}\label{app} In this Appendix we show the nonparametric estimator (\ref{eqnpR}) is nonnegative definite, from which it follows that (\ref{eqnp}) and (\ref{eqnpmissing}) are also nonnegative definite. Below, we present an argument that (\ref{eqnp}) is asymptotically unbiased for $C_{ij}(\mathbf{x},\by,d(t_0))$. The nonnegative definiteness property is not restricted to a bivariate process, so assume there are $p$ spatial processes $W_i(\bs,t), i=1,\ldots,p$, with observation network locations $\bs_m, m=1,\ldots,n$. Define $U_i(\bs,t) = \break W_i(\bs,t) \indicator_{[W_i(\mathbf{s},t)\ \mathrm{is\ observed}]}$, noting that if $W_i(\bs,t)$ is unavailable at a particular location and time, $U_i(\bs,t) = 0$. Consider evaluating $R_{ij}(\mathbf{x}_k,\mathbf{x}_\ell,t)$ at any arbitrary locations $\mathbf{x}_k$ and $\mathbf{x}_\ell,\break k,\ell=1,\ldots,N$, and define the arbitrary vector $\ba= (a_{11},\ldots,a_{1N},a_{21},\ldots,a_{pN}).$ Set $\Sigma$ to be the covariance matrix made up of the functions $R_{ij}(\cdot,\cdot,t)$ corresponding to the random vector \[ \bigl(W_1(\mathbf{x}_1,t),W_1( \mathbf{x}_2,t),\ldots,W_p(\mathbf{x}_N,t) \bigr)'. \] Then, absorbing the denominator into the kernel functions of $R_{ij}(\mathbf{x}_k,\mathbf{x}_\ell,t)$, and writing $R_{ij}' (\mathbf{x}_k,\mathbf{x}_\ell,t)$ for this normalized function, we have \begin{eqnarray*} \ba' \Sigma\ba&=& \sum_{i,j=1}^p \sum_{k,\ell=1}^N a_{ik} a_{j\ell} R_{ji}'(\mathbf{x}_\ell, \mathbf{x}_k,t) \\ &= &\sum_{i,j=1}^p \sum _{k,\ell=1}^N a_{ik} a_{j\ell} \sum _{m,r=1}^n K_\lambda\bigl(\| \mathbf{x}_\ell- \bs_m\|\bigr) K_\lambda\bigl(\|\mathbf{x}_k - \bs_r\|\bigr) U_i(\bs_r,t) U_j( \bs_m,t) \\ &= &\sum_{m,r=1}^n \sum _{i,j=1}^p \sum_{k,\ell=1}^N \bigl(a_{ik} K_\lambda\bigl(\|\mathbf{x}_k - \bs_r \|\bigr) U_i(\bs_r,t) \bigr) \bigl(a_{j\ell} K_\lambda\bigl(\|\mathbf{x}_\ell- \bs_m\|\bigr) U_j( \bs_m,t) \bigr) \\ &= &\Biggl( \sum_{r=1}^n \sum _{i=1}^p \sum_{k=1}^N a_{ik} K_\lambda\bigl(\|\mathbf{x}_k - \bs_r\|\bigr) U_i(\bs_r,t) \Biggr)^2 \geq0. \end{eqnarray*} To show that (\ref{eqnp}) is asymptotically unbiased for $C_{ij}(\mathbf{x},\by,d(t_0))$, we disregard the smoothing over time, since asymptotically we do not have a finer resolution of time points (but for consistency we would assume an increasing number of realizations per each day of the year). In particular, suppose we observe the bivariate process $(W_N(\bs _k),W_X(\bs_k))$ for $\bs_1,\ldots,\bs_n\in\mathcal{D} \subset\real^d$, which are samples from a distribution with strictly positive probability density $f\dvtx\mathcal{D}\rightarrow\real^+$, with empirical c.d.f. $F_n(\mathbf{x}) = \frac{1}{n}\sum_{k=1}^n \indicator_{[\mathbf{s}_k \leq \mathbf{x} ]}$, where the indicator function is 1 if the inequality holds for all indices of $\mathbf{x}$. The density~$f$, with corresponding c.d.f. $F$, allows the network density to vary across the domain. We additionally suppose $n\rightarrow\infty$ and $\lambda\rightarrow0$ such that $\lambda\sim n^{-1/d + \varepsilon}$ for some small $0<\varepsilon<1/d^2$. Suppressing the time indexing from our notation, we can write \[ \hat{C}_{ij}(\mathbf{x},\by) = \frac {1}{n^2\lambda^{2d}}\sum _{k=1}^n \sum_{\ell=1}^n K_\lambda' \bigl(\|\mathbf{x}-\bs_k\| \bigr) K_\lambda' \bigl(\|\by-\bs_\ell\| \bigr) W_i(\bs_k) W_j(\bs_\ell), \] where the denominator of (\ref{eqnp}) is absorbed into the kernel functions of the numerator, yielding standardized functions $K_\lambda'$. Here we only consider the direct covariance estimators at a location $\mathbf{x}\in{\cal D}\setminus\partial{\cal D}$, $C_{ii}(\mathbf{x},\mathbf{x})$; the same argument applies for the direct and cross-covariance functions $C_{ij}(\mathbf{x},\by)$ for $\mathbf{x}\not=\by$. We have \begin{eqnarray} \E\hat{C}_{ii}(\mathbf{x},\mathbf{x}) &=& \frac{1}{n^2 \lambda^{2d}} \sum _{k,\ell=1}^n K_\lambda'\bigl (\|\mathbf{x}- \bs_k\| \bigr)K_\lambda' \bigl(\|\mathbf{x}-\bs_\ell\| \bigr) \E\bigl(W_i(\bs_k) W_i(\bs_\ell) \bigr) \\ &=& \frac{1}{n^2 \lambda^{2d}} \sum_{k,\ell=1}^n K_\lambda' \bigl(\|\mathbf{x}-\bs_k\| \bigr)K_\lambda' \bigl(\|\mathbf{x}-\bs_\ell\| \bigr) C_{ii}(\bs_k, \bs_\ell) \\ &&{} + \frac{1}{n^2 \lambda^{2d}} \sum_{k = 1}^n K_\lambda' \bigl(\|\mathbf{x}-\bs_k\| \bigr)^2 \tau_i(\bs_k,\bs_k)^2. \end{eqnarray} Invoking Lemma 7 of \citet{kleiber2012nsmm}, in the limit as $n\rightarrow\infty$, we can pass from the sum to the integral. Assume the empirical c.d.f. $F_n$ is close to the limiting c.d.f. $F$, where $\sup_{\mathbf{x}} |F_n(\mathbf{x}) - F(\mathbf{x})| = D_n$ where $D_n = o(1/(n\lambda^d))$. This rate holds, for example, if ${\cal D} = [0,1]$, $F$ is the uniform density and $F_n$ is the empirical c.d.f. of the uniform grid $(1/n,2/n,\ldots,n/n)$. For $d > 1$, if $n$ grows as $M^d$, a rate of $D_n \sim1/n^{1/d}$ can be derived for sampling locations on a regular grid with limiting uniform distribution [\citet{kleiber2012nsmm}]. Then we have \begin{eqnarray} \E\hat{C}_{ii}(\mathbf{x},\mathbf{x}) &= &\frac{1}{\lambda^{2d}} \iint_{{\cal D}^2} K_\lambda' \bigl(\|\bu-\mathbf{x}\| \bigr)K_\lambda' \bigl(\| \bv-\mathbf{x}\| \bigr) C_{ii}(\bu,\bv) \,\mathrm{ d}F(\bu )\,\mathrm{ d}F(\bv) \nonumber \\[-6pt] \\[-10pt] \nonumber & &{}+ \frac{1}{n \lambda^{2d}} \int_{{\cal D}} K_\lambda' \bigl(\|\bu-\mathbf{x}\| \bigr)^2 \tau_i(\bu,\bu)^2 \,\mathrm{ d}F( \bu) + {\cal O}(D_n). \end{eqnarray} Making the change of variables to $\ba= (\bu- \mathbf{x})/\lambda$ and $\bb= (\bv- \mathbf{x})/\lambda$ yields \begin{eqnarray} \label{eqasymptoticbias} &&\iint_{{\cal D'}^2} K'\bigl (\|\ba\| \bigr)K'\bigl (\|\bb\| \bigr) C_{ii}(\lambda\ba+ \mathbf{x},\lambda\bb+ \mathbf{x}) { \,\mathrm d}F(\ba) \,\mathrm{ d}F(\bb) \nonumber \\[-8pt] \\[-8pt] \nonumber &&\qquad{} + \frac{1}{n \lambda^{d}} \int_{{\cal D'}} K' \bigl(\|\ba\| \bigr)^2 \tau_{i}(\lambda\ba+ \mathbf{x},\lambda\ba+ \mathbf{x})^2\,\mathrm{ d}F(\ba) + {\cal O}(D_n) \end{eqnarray} for an appropriate translated domain ${\cal D'}$. As $\lambda\sim n^{-1/d + \varepsilon}$, the second term of (\ref{eqasymptoticbias}) converges to zero. The arguments from \citet{kleiber2012nsmm} applied to the first term of (\ref{eqasymptoticbias}) then yield the unbiasedness of $C_{ii}(\mathbf{x},\mathbf{x})$. \end{appendix}
\section{Introduction} \seclabel{intro} A \emph{loop} $(Q,\cdot)$ is a set $Q$ with a binary operation $\cdot : Q\times Q\to Q$ such that (i) $(Q,\cdot)$ is a \emph{quasigroup}, that is, for each $a$, $b\in Q$, the equations $ax=b$ and $ya=b$ have unique solutions $x$, $y\in Q$, and (ii) there exists a neutral element $1\in Q$ such that $1x = x1 = x$ for all $x\in Q$. Equivalently, a loop can be viewed as having three binary operations $\cdot, \ldiv, \rdiv$ satisfying the identities $x\ldiv (xy) = y$, $x(x\ldiv y) = y$, $(xy)\rdiv y = x$, $(x\rdiv y)y = x$, $x\rdiv x = y\ldiv y$. Basic references for loop theory are \cite{Bruck, Pflugfelder}. For $a\in Q$, the \emph{right translation} and \emph{left translation} by $a$ are the bijections $R_a :Q\to Q; x\mapsto xa$ and $L_a :Q\to Q;x\mapsto ax$. These generate the \emph{multiplication group} $\Mlt(Q) = \sbl{R_x, L_x\ |\ x\in Q}$. The \emph{inner mapping group} is the subgroup stabilizing the neutral element, $\Inn(Q) = (\Mlt(Q))_1$. A loop is \emph{automorphic} (or an \emph{A-loop}) if every inner mapping is an automorphism, that is, if $\Inn(Q) \leq \Aut(Q)$. The study of automorphic loops began in 1956 with Bruck and Paige \cite{BP}. They were particularly interested in \emph{diassociative} automorphic loops, that is, loops in which each $2$-generated subloop is a group. They noted that such loops share many properties with Moufang loops. Shortly thereafter, Osborn showed that commutative diassociative automorphic loops are Moufang \cite{Osborn}. More results showing the Moufang nature of diassociative, automorphic loops were found in \cite{Drapal} and \cite[Thm. 5]{Wright1}. The general case was finally settled in \cite{KKP}: Every diassociative automorphic loop is a Moufang loop. In recent years, a detailed structure theory has emerged for \emph{commutative} automorphic loops. For instance, the Odd Order, Lagrange and Cauchy Theorems hold for commutative automorphic loops, a finite commutative automorphic loop has order $p^k$ if and only if each element has order a power of $p$, and a finite commutative automorphic loop decomposes as a direct product of a loop of odd order and a loop of order a power of $2$ \cite{JKV1}; there are no finite simple nonassociative commutative automorphic loops \cite{GKN}; for an odd prime $p$, if $Q$ is a finite commutative automorphic $p$-loop then $\Mlt(Q)$ is a $p$-group and $Q$ is centrally nilpotent \cite{Csorgo1,JKV3}; for an odd prime $p$, commutative automorphic loops of order $p$, $p^2$, $2p$, $2p^2$, $4p$, $4p^2$ are groups \cite{JKV2}. In this paper we lay foundations for the study of automorphic loops. Our understanding is not yet as complete as in the commutative case, but we obtain several significant results, as described below. For notation and terminology, see Section \secref{preliminaries}. \subsection{Summary of results} \S \secref{preliminaries} introduces the notation, definitions, and preliminary results concerned mostly with identities valid in automorphic loops. \S\S \secref{cores}, \secref{bruck}: Motivated by work of Glauberman, we first study certain derived operations on automorphic loops. In \cite{Gl1,Gl2} Glauberman showed that Bruck loops of odd order are solvable and satisfy the Cauchy, Lagrange and Sylow Theorems. He also constructed a Bruck loop $(Q,\circ)$ from a uniquely $2$-divisible Moufang loop $(Q,\cdot)$ by setting $x\circ y = (xy^2x)^{1/2}$, and this allowed him to transfer the above results from Bruck loops of odd order to Moufang loops of odd order. We show in three steps that the analog of Glauberman's operation for uniquely $2$-divisible automorphic loops is the operation \begin{displaymath} x\circ y = (x^{-1}\ldiv (y^2x))^{1/2}, \end{displaymath} which coincides with Glauberman's operation on Moufang loops because Moufang loops are diassociative. First, given any automorphic loop $(Q,\cdot)$, we show that the core $(Q,*)$ defined by \begin{displaymath} x*y = x^{-1}\ldiv (y^{-1}x). \end{displaymath} is an involutive quandle. Second, using the core, we show that the set $P_Q = \{P_x\ |\ x\in Q\}$ with $P_x = R_xL_{x^{-1}}^{-1}$ is a twisted subgroup of $\Mlt(Q)$, satisfying $P_xP_yP_y = P_{yP_x}$ and $P_x^n = P_{x^n}$. As is well known, on a uniquely $2$-divisible twisted subgroup $(T,\cdot)$ one can define a Bruck loop $(T,\bullet)$ by \begin{displaymath} x\bullet y = (xy^2x)^{1/2}. \end{displaymath} Hence, if $Q$ is uniquely $2$-divisible, $(P_Q,\bullet)$ is a twisted subgroup. Third, the operation $\bullet$ can be transferred from $P_Q$ onto $Q$, yielding the associated Bruck loop $(Q,\circ)$. A finite automorphic loop is uniquely $2$-divisible if and only if it is of odd order. The above discussion therefore applies to automorphic loops of odd order, and then results of Glauberman on Bruck loops lead to the Lagrange and Cauchy (but not Sylow) Theorems for automorphic loops of odd order. \S\S \secref{correspondence}, \secref{solvability}: The next ingredient is based on Wright's construction of loops from algebras. Specializing it to a Lie ring $(Q,+,[\cdot,\cdot])$, we can define $(Q,\diamond)$ by \begin{displaymath} x\diamond y = x+y-[x,y]. \end{displaymath} Then $(Q,\diamond)$ is a loop if and only if in $(Q,+,[\cdot,\cdot])$ the mappings \begin{equation}\label{Eq:Wright} y\mapsto y\pm [y,x]\text{ are invertible for every }x\in Q.\tag{$W_1$} \end{equation} Moreover, if $(Q,+,[\cdot,\cdot])$ is a Lie ring satisfying \eqref{Eq:Wright}, then a sufficient condition for $(Q,\diamond)$ to be automorphic is that \begin{equation}\label{Eq:WrightAutomorphic} [[Q,x],[Q,x]]=1\text{ for every $x\in Q$.}\tag{$W_2$} \end{equation} In the uniquely $2$-divisible case we obtain the following correspondence: If $(Q,+,[\cdot,\cdot])$ is a uniquely $2$-divisible Lie ring satisfying \eqref{Eq:Wright} and \eqref{Eq:WrightAutomorphic}, then $(Q,\diamond)$ is a uniquely $2$-divisible automorphic loop whose associated Bruck loop $(Q,\circ)$ is an abelian group. Conversely, if $(Q,\cdot)$ is a uniquely $2$-divisible automorphic loop whose associated Bruck loop $(Q,\circ)$ is an abelian group, then $(Q,\circ,[\cdot,\cdot])$ defined by \begin{displaymath} [x,y]=x\circ y\circ (xy)^{-1} \end{displaymath} (the inverses in $(Q,\cdot)$ and $(Q,\circ)$ coincide) is a Lie ring satisfying \eqref{Eq:Wright} and \eqref{Eq:WrightAutomorphic}. Moreover, the two constructions are inverse to each other, subrings (resp. ideals) of the Lie ring are subloops (resp. normal subloops) of the automorphic loop, and subloops (resp. normal subloops) of the automorphic loop closed under square roots are subrings (resp. ideals) of the Lie ring. Taking advantage of the associated Lie rings, we prove the Odd Order Theorem for automorphic loops, we show that automorphic loops of order $p^2$ are groups, and we give examples of automorphic loops of order $p^3$ with trivial nucleus. \S \secref{simplicity}: Next we investigate finite simple automorphic loops. Since a loop $Q$ is simple if and only if $\Mlt(Q)$ is a primitive permutation group on $Q$, we approach the problem from the direction of primitive groups. In \cite{JKNV} we proved computationally, using the library of primitive groups in GAP, that a finite simple automorphic loop of order less than $2500$ is associative. Here we show that if $Q$ is a finite simple nonassociative automorphic loop then the socle of $\Mlt(Q)$ is not regular, hence, by the O'Nan-Scott theorem, $\Mlt(Q)$ is of almost simple type, of diagonal type or of product type. Whether such a loop exists remains open. We also prove that characteristically simple automorphic loops behave analogously to characteristically simple groups. \S\S \secref{mid_nuc}, \secref{dihedral}: We conclude the paper with a short discussion of middle nuclear extensions and, as an application, with constructions of generalized dihedral automorphic loops. Namely, if $(A,+)$ is an abelian group and $\alpha\in\mathrm{Aut}(A)$ then $\mathbb Z_2\times A$ with multiplication $(i,u)(j,v) = (i + j, ((-1)^ju + v)\alpha^{ij})$ is an automorphic loop. In particular, if $A=\mathbb Z_n$ and $c$ is an invertible element of $\mathbb Z_n$, then $\mathbb Z_2\times \mathbb Z_n$ with multiplication $(i,u)(j,v) = (i + j, ((-1)^ju + v)c^{ij})$ is a dihedral automorphic loop. We show that two such loops are isomorphic if and only if the invertible elements coincide, and we calculate the automorphism groups of these loops. Cs\"org\H{o} showed in \cite{Csorgo2} that if $Q$ is a finite automorphic loop and $x\in Q$ then $|x|$ divides $|Q|$. This allows us to classify all automorphic loops of order $2p$. There are $p$ such loops up to isomorphism; these are precisely the dihedral automorphic loops corresponding to the $p-1$ invertible elements of $\mathbb Z_p$, and the cyclic group $\mathbb Z_{2p}$. \section{Preliminaries} \seclabel{preliminaries} The inner mapping group $\Inn(Q)$ has a standard set of generators \cite{Bruck}: \[ R_{x,y} = R_x R_y R_{xy}\inv\,, \qquad T_x = R_x L_x\inv\,, \qquad L_{x,y} = L_x L_y L_{yx}\inv\,. \] Thus automorphic loops can be characterized equationally. \begin{proposition}[\cite{BP}] \prplabel{basic} A loop $Q$ is an automorphic loop if and only if, for all $x$, $y$, $u$, $v\in Q$, \begin{align*} (uv)R_{x,y} &= uR_{x,y}\cdot vR_{x,y}, \tag{$A_r$} \\ (uv)L_{x,y} &= uL_{x,y}\cdot vL_{x,y}, \tag{$A_l$} \\ (uv)T_x &= uT_x\cdot vT_x. \tag{$A_m$} \end{align*} \end{proposition} This means that automorphic loops form a variety in the sense of universal algebra. In particular, subloops and factor loops of automorphic loops are automorphic \cite[Thm. 2.2]{BP}. A loop $Q$ is \emph{power-associative} if for each $x\in Q$, $\sbl{x}$ is a group. In particular, powers of $x$ are unambiguous, and $x^m x^n = x^{m+n}$ for all $m$, $n\in \mathbb{Z}$. \begin{proposition} [\text{\cite[Thm. 2.4]{BP}}] \prplabel{power-associative} Every automorphic loop is power-associative. \end{proposition} We will use the power-associativity of automorphic loops without explicitly referring to Proposition \prpref{power-associative}. \begin{proposition}[\text{\cite[Thm. 2.5]{BP}}] \prplabel{commutes} Let $Q$ be an automorphic loop. Then the following hold for all $x\in Q$, $j$, $k$, $m$, $n\in \mathbb{Z}$. \begin{align} L_{x^m}^j L_{x^n}^k &= L_{x^n}^k L_{x^m}^j, \eqnlabel{LLLL} \\ R_{x^m}^j L_{x^n}^k &= L_{x^n}^k R_{x^m}^j, \eqnlabel{RLLR} \\ R_{x^m}^j R_{x^n}^k &= R_{x^n}^k R_{x^m}^j. \eqnlabel{RRRR} \end{align} \end{proposition} \begin{corollary} \corlabel{RLxxRLxx} For all $x$ in an automorphic loop $Q$, \begin{align} L_{x,x\inv} &= L_{x\inv,x}, \eqnlabel{LxxLxx} \\ R_{x,x\inv} &= R_{x\inv,x}. \eqnlabel{RxxRxx} \end{align} \end{corollary} A loop $Q$ is said to have the \emph{antiautomorphic inverse property} (AAIP) if it has two-sided inverses and satisfies the identity \[ (xy)\inv = y\inv x\inv \tag{AAIP} \] for all $x$, $y\in Q$. It is also useful to characterize the AAIP in terms of translations and the \emph{inversion mapping} $J: Q\to Q; x\mapsto x\inv$ as either of the following: \begin{align} R_x^J &= L_{x\inv}, \eqnlabel{aaip1} \\ L_x^J &= R_{x\inv}. \eqnlabel{aaip2} \end{align} \begin{proposition}[\text{\cite[Coro. 6.6]{JKNV}}] \prplabel{AAIP} Every automorphic loop has the AAIP. \end{proposition} \begin{corollary} \corlabel{AAIP} If $Q$ is an automorphic loop, then $J$ normalizes $\Mlt(Q)$ in $\Sym(Q)$. \end{corollary} \begin{proof} Since $\Mlt(Q)$ is generated by left translations, this follows from \eqnref{aaip1} and \eqnref{aaip2} in view of Proposition \prpref{AAIP} \end{proof} \begin{lemma} \lemlabel{RxyLxy} In an automorphic loop $Q$, the following hold for all $x$, $y\in Q$. \begin{align} R_{x,y} &= L_{x\inv,y\inv}, \eqnlabel{RxyLxy} \\ T_x\inv &= T_{x\inv}. \eqnlabel{Tinv} \end{align} \end{lemma} \begin{proof} We compute \[ R_{x,y} = R_{x,y}^J = R_x^J R_y^J (R_{xy}\inv)^J = L_{x\inv} L_{y\inv} L_{(xy)\inv}\inv = L_{x\inv} L_{y\inv} L_{y\inv x\inv}\inv = L_{x\inv,y\inv}\,, \] where we used $R_{x,y}\in \Aut(Q)$ in the first equality, \eqnref{aaip1} in the third, and (AAIP) in the fourth. This establishes \eqnref{RxyLxy}. For \eqnref{Tinv}, we have \[ T_x T_{x\inv} = R_x L_x\inv R_{x\inv} L_{x\inv}\inv = R_x R_{x\inv} L_x\inv L_{x\inv}\inv = R_{x,x\inv} L_{x\inv,x}\inv = R_{x,x\inv} R_{x,x\inv}\inv = \id_Q\,, \] where we used \eqnref{RLLR} in the second equality, and \eqnref{RxyLxy} in the fourth. \end{proof} To check that a particular loop is automorphic, it is not necessary to verify all of the conditions ($A_r$), ($A_{\ell}$) and ($A_m$): \begin{proposition}[\text{\cite[Thm. 6.7]{JKNV}}] \prplabel{halfcheck} Let $Q$ be a loop satisfying $(A_m)$ and $(A_{\ell})$. Then $Q$ is automorphic. \end{proposition} The \emph{left}, \emph{right}, and \emph{middle nucleus} of a loop $Q$ are defined, respectively, by \begin{align*} N_{\lambda}(Q) &= \{ a\in Q\mid ax\cdot y = a\cdot xy,\quad \forall x,\,y\in Q\}, \\ N_{\rho}(Q) &= \{ a\in Q\mid xy\cdot a = x\cdot ya,\quad \forall x,\,y\in Q\}, \\ N_{\mu}(Q) &= \{ a\in Q\mid xa\cdot y = x\cdot ay,\quad \forall x,\,y\in Q\}, \end{align*} and the \emph{nucleus} is $N(Q) = N_{\lambda}(Q)\cap N_{\rho}(Q)\cap N_{\mu}(Q)$. Each of these is a subloop. Recall that a subloop $S\le Q$ is normal in $Q$, $S\unlhd Q$, if $(S)\varphi = S$ for all $\varphi\in\Inn(Q)$. \begin{proposition} \prplabel{nuclei} Let $Q$ be an automorphic loop. Then \begin{enumerate} \item[(i)] $N_{\lambda}(Q) = N_{\rho}(Q) \subseteq N_{\mu}(Q)$, and \item[(ii)] each nucleus is normal in $Q$. \end{enumerate} \end{proposition} \begin{proof} The equality $N_{\lambda}(Q) = N_{\rho}(Q)$ is an immediate consequence of the AAIP. Suppose $a \in N_{\lambda}(Q)$. Then $a\inv \in N_{\lambda}(Q)$ and $(x)T_a = a\inv x a$. Now for all $x$, $y\in Q$, \[ (x\cdot ay)T_a = (x)T_a\cdot (ay)T_a = (a\inv x a)\cdot ya = a\inv (xa\cdot y) a = (xa\cdot y)T_a\,, \] where we used ($A_m$) in the first equality, and the equality of the left and right nuclei in the third. Since $T_a$ is a permutation, we have $x\cdot ay = xa\cdot y$ for all $x$, $y\in Q$, that is, $a\in N_{\mu}(Q)$. This establishes (i). Part (ii) is \cite[Thm. 2.2(iii)]{BP}. \end{proof} For a subset $S$ of a loop $Q$, we define the \emph{commutant of} $S$ to be the set \[ C_Q(S) = \{ a\in Q\mid ax = xa\quad \text{for all}\quad x\in S \}\,. \] The \emph{commutant} of $Q$ itself, $C_Q(Q)$ is just denoted by $C(Q)$. (In a group, the commutant of a set is the centralizer of the set and the commutant is the center. However, ``center'' has a narrower meaning in loop theory, and so we adapt operator theory terminology to the present setting.) \begin{proposition} \prplabel{commutant} Let $Q$ be an automorphic loop and let $S\subseteq Q$. Then $C_Q(S) \leq Q$. Furthermore, if $S\normal Q$ then $C_Q(S)\normal Q$. In particular, the commutant $C(Q)$ is a normal subloop of $Q$. \end{proposition} \begin{proof} We have $a\in C_Q(S)$ if and only if $(a)T_x = a$ for all $x\in S$. Thus $C_Q(S)$ is characterized as the intersection of the fixed point sets of all $T_x$, $x\in S$. Since $T_x\in\Aut(Q)$, the fixed point set of $T_x$ is a subloop of $Q$, and $C_Q(S)\le Q$ follows. Now suppose $S\normal Q$. Fix $a\in C_Q(S)$, $x\in S$, $\varphi\in \Inn(Q)$ and set $y = (x)\varphi\inv\in S$. Then \[ x(a)\varphi = (y)\varphi(a)\varphi = (ya)\varphi = (ay)\varphi = (a)\varphi(y)\varphi = (a)\varphi x\,, \] using $\varphi\in \Aut(Q)$ in the first and fourth equalities and $a\in C_Q(S)$ in the third. Since $x\in S$ was arbitrary, $(a)\varphi\in C_Q(S)$. Thus $C_Q(S)\unlhd Q$. \end{proof} We conclude the section with several definitions needed throughout the paper. A subset $S$ of a loop $Q$ is said to be \emph{characteristic} in $Q$, denoted by $S\Char Q$, if for every $\varphi\in \Aut(Q)$, $(S)\varphi = S$. A loop is \emph{characteristically simple} if it has no nontrivial characteristic subloops. A loop is \emph{simple} if it has no nontrivial normal subloops. A loop $Q$ is \emph{solvable} if it has a subnormal series $1= Q_0 \leq \cdots \leq Q_n = Q$, $Q_i \normal Q_{i+1}$, such that each factor loop $Q_{i+1}/Q_i$ is an abelian group. The \emph{derived subloop} $Q'$ of a loop $Q$ is the smallest normal subloop of $Q$ such that $Q/Q'$ is an abelian group. The derived subloop can be characterized as the smallest normal subloop containing each \emph{commutator} $[x,y]$, defined by $xy\cdot [y,x]=yx$, and each \emph{associator} $[x,y,z]$, defined by $xy\cdot z = (x\cdot yz)[x,y,z]$. Since automorphisms evidently map commutators to commutators and associators to associators, it follows that $Q' \Char Q$. The \emph{higher derived subloops} are defined in the usual way: $Q^{(2)} = Q'' = (Q')'$, $Q^{(3)} = Q'''$, \emph{etc}. Note that a loop $Q$ is solvable if and only if $Q^{(n)} = 1$ for some $n > 0$. A \emph{Bruck loop} is a loop satisfying the \emph{left Bol identity} $(x(yx))z = x(y(xz))$ and the \emph{automorphic inverse property} $(xy)\inv = x\inv y\inv$. \section{Cores and twisted subgroups} \seclabel{cores} In an automorphic loop $Q$, we introduce a new binary operation $\ast$ as follows: \[ x\ast y = x\inv \ldiv (y\inv x) = (x\inv \ldiv y\inv) x \tag{$\ast$} \] for all $x$, $y\in Q$. (The second equality follows from \eqnref{RLLR}.) We will refer to the magma $(Q,\ast)$ as the \emph{core} of the loop $Q$, which should not be confused with the core of a subgroup in group theory. A similar notion was introduced by Bruck \cite{Bruck} for Moufang loops (where the operation can be more simply written as $xy\inv x$) and also in our previous papers \cite{JKV1,JKV2} in the commutative case. As in \cite{Gl2,JKV1,JKV2}, it is useful to introduce the following permutations for each $x$ in an automorphic loop $Q$: \[ P_x = R_x L_{x\inv}^{\inv} = L_{x^{-1}}^{-1} R_x, \tag{P} \] where the second equality follows by \prpref{commutes}. Thus the left translation maps of the core $(Q,\ast)$ are just the maps $J P_x$, $x\in Q$; a fact we will use heavily. \begin{proposition} \prplabel{coreaut} Let $Q$ be an automorphic loop with core $(Q,\ast)$. Then for all $x$, $y$, $z\in Q$, \begin{align} (y\ast z)R_x &= yR_x\ast zR_x, \eqnlabel{Rcoreaut}\\ (y\ast z)L_x &= yL_x\ast zL_x. \eqnlabel{Lcoreaut} \end{align} Therefore $\Mlt(Q) \leq \Aut(Q,\ast)$. In particular, $P_x\in\Aut(Q,*)$ for all $x\in Q$. \end{proposition} \begin{proof} We start with \eqnref{RxyLxy}, which we write as $R_{y,x} = L_{y\inv,x\inv}$, that is, $L_{y\inv} L_{x\inv} L_{(yx)\inv}\inv = R_y R_x R_{yx}\inv$. Rearranging this, we have $L_{x\inv} L_{(yx)\inv}\inv R_{yx} = L_{y\inv}\inv R_y R_x$, or \begin{equation} \eqnlabel{PLnorm} L_{x\inv} P_{yx} = P_y R_x\,. \end{equation} Applying both sides of \eqnref{PLnorm} to $z\inv$ yields $(yx)\inv \ldiv [(x\inv z\inv)\cdot yx] = [y\inv \ldiv (z\inv y)]x$. Since $x\inv z\inv = (zx)\inv$ by the AAIP, we have \eqnref{Rcoreaut}. To establish \eqnref{Lcoreaut}, observe first that $((1\rdiv y)x^{-1})^{-1} = x(1\rdiv y)^{-1} = xy$ by AAIP, and so $R_y\inv R_{x\inv} R_{xy}$ is an inner mapping, hence an automorphism. Thus \[ R_y\inv R_{x\inv} R_{xy} = (R_y\inv R_{x\inv} R_{xy})^J = (R_y\inv)^J R_{x\inv}^J R_{xy}^J = L_{y\inv}\inv L_x L_{(xy)\inv}\,, \] using \eqnref{aaip1} and \eqnref{aaip2}. Rearranging, we have $R_{x\inv} R_{xy} L_{(xy)\inv}\inv = R_y L_{y\inv}\inv L_x$, or \begin{equation} \eqnlabel{PRnorm} R_{x\inv} P_{xy} = P_y L_x\,. \end{equation} Applying both sides of \eqnref{PRnorm} to $z\inv$ yields $(xy)\inv \ldiv [(z\inv x\inv)\cdot xy] = x[y\inv \ldiv (z\inv y)]$. Since $z\inv x\inv = (xz)\inv$ by the AAIP, we are finished. \end{proof} \begin{lemma} For all $x$ in an automorphic loop $Q$, \begin{equation} \eqnlabel{Pinv} P_x^J = P_x^{\inv} = P_{x\inv}\,. \end{equation} Thus in the core $(Q,\ast)$, the following holds for all $x$, $y\in Q$: \begin{equation} \eqnlabel{coreinv} (x \ast y)\inv = x\inv \ast y\inv\,. \end{equation} \end{lemma} \begin{proof} We have $P_x^J = R_x^J (L_{x\inv}\inv)^J = L_{x\inv} R_x\inv = P_x\inv$, using \eqnref{aaip1} and \eqnref{aaip2}. Also, \[ P_x P_{x\inv} = R_x L_{x\inv}\inv R_{x\inv} L_x\inv = R_x L_x\inv R_{x\inv} L_{x\inv}\inv = T_x T_{x\inv} = \id_Q\,, \] using \eqnref{RLLR} and \eqnref{LLLL} in the second equality and \eqnref{Tinv} in the fourth. This establishes \eqnref{Pinv}. Then \eqnref{coreinv} follows, since $(x\ast y)\inv = yJP_xJ = yP_x^J = (y\inv)JP_{x\inv} = x\inv \ast y\inv$. \end{proof} \begin{theorem} \thmlabel{coreprops} Let $Q$ be an automorphic loop with core $(Q,\ast)$. Then $(Q,\ast)$ is an involutive quandle, that is, the following properties hold: \begin{enumerate} \item[(i)] $x\ast x = x$ for all $x\in Q$, \item[(ii)] $x\ast (x\ast y) = y$ for all $x$, $y\in Q$, \item[(iii)] $x\ast (y\ast z) = (x\ast y)\ast (x\ast z)$ for all $x$, $y$, $z\in Q$. \end{enumerate} \end{theorem} \begin{proof} Part (i) is clear from the definition of $\ast$. For (ii), $x\ast (x\ast y) = yJP_xJP_x = yP_x^J P_x = y$ by \eqnref{Pinv}. For (3), \[ x\ast (y\ast z) = (y\ast z)JP_x = (y\inv \ast z\inv)P_x = (y\inv)P_x \ast (z\inv)P_x = (x\ast y)\ast (x\ast z)\,, \] using \eqnref{coreinv} and Proposition \prpref{coreaut}. \end{proof} Recall that a subset $A$ of a group $G$ is said to be a \emph{twisted subgroup} of $G$ if (i) $1\in A$, (ii) $a\in A$ implies $a\inv \in A$, and (iii) $a,b\in A$ implies $aba\in A$. In an automorphic loop $Q$, let $P_Q = \setof{P_x}{x\in Q}$. \begin{proposition} \prplabel{twisted} Let $Q$ be an automorphic loop. Then $P_Q$ is a twisted subgroup of $\Mlt(Q)$. In particular, \begin{equation} \eqnlabel{Ptwisted} P_x P_y P_x = P_{yP_x} \end{equation} for all $x$, $y\in Q$. \end{proposition} \begin{proof} Clearly $\id_Q = P_1\in P_Q$. For $x\in Q$, $P_x\inv \in P_Q$ by \eqnref{Pinv}. Since $JP_x \in \Aut(Q,\ast)$ by Theorem \thmref{coreprops}(iii), we have $zJP_y JP_x = (y\ast z)JP_x = yJP_x\ast zJP_x = zJP_x JP_{yJP_x}$ for all $x$, $y$, $z\in Q$. Thus $P_y^J P_x = P_x^J P_{(y\inv)P_x}$. By \eqnref{Pinv}, we deduce $P_x P_{y\inv} P_x = P_{(y\inv)P_x}$. Replacing $y$ with $y\inv$, we have \eqnref{Ptwisted}. \end{proof} \begin{corollary} \corlabel{Ppowers} Let $Q$ be an automorphic loop. Then for all $x\in Q$ and $n\in \mathbb{Z}$, \begin{equation} \eqnlabel{Ppowers} P_x^n = P_{x^n}\,. \end{equation} \end{corollary} \begin{proof} Since $(x^n)P_x = x^{n+2}$, the desired result follows for $n\geq 0$ by an easy induction using \eqnref{Ptwisted}. For $n < 0$, apply \eqnref{Pinv}. \end{proof} Although we have no application for the following result, we mention it for the sake of completeness: \begin{proposition} \prplabel{Pnormal} Let $Q$ be an automorphic loop. Then $\langle P_Q\rangle \normal \Mlt(Q)$. \end{proposition} \begin{proof} By \eqnref{PLnorm}, we have for each $x,y\in Q$, $R_x\inv P_y R_x = R_x\inv L_{x\inv} P_{yx} = P_x\inv P_{yx} \in \langle P_Q\rangle$. By \eqnref{PRnorm}, we have for each $x,y\in Q$, $L_x\inv P_y L_x = L_x\inv R_{x\inv} P_{xy} = P_{x\inv} P_{xy}\in \langle P_Q\rangle$. Since $\Mlt(Q)$ is generated by all $R_x$, $L_x$, $x\in Q$, we have the desired result. \end{proof} \section{Uniquely $2$-divisible automorphic loops} \seclabel{bruck} A loop $Q$ is said to be \emph{uniquely} $2$-\emph{divisible} if the squaring map $x\mapsto x^2$ is a permutation of $Q$. \begin{lemma} \lemlabel{Pcorrespond} Let $Q$ be a uniquely $2$-divisible automorphic loop. Then $Q\to P_Q; x\mapsto P_x$ is a bijection. \end{lemma} \begin{proof} To see that the map is one-to-one, suppose $P_x = P_y$. Applying both sides to $1$, we obtain $x^2 = y^2$. By unique $2$-divisibility, $x = y$. \end{proof} It is well known that a uniquely $2$-divisible twisted subgroup $T$ of a group $G$ can be turned into a Bruck loop $(T,\bullet)$ by setting \begin{displaymath} a\bullet b = (ab^2a)^{1/2}\tag{$\bullet$}. \end{displaymath} See \cite[Lem. 4.5]{FKP}, for instance. In a uniquely $2$-divisible automorphic loop $Q$, the set $P_Q$ is a uniquely $2$-divisible twisted subgroup of $\Mlt(Q)$ by Proposition \prpref{twisted} and Corollary \corref{Ppowers}, noticing that $P_x^{1/2} = P_{x^{1/2}}$ for all $x\in Q$. Thus we can define \[ P_x\bullet P_y = [P_x P_y^2 P_x]^{1/2} = P_{(y^2)P_x}^{1/2} = P_{[(y^2)P_x]^{1/2}}, \] making $(P_Q,\bullet)$ into a Bruck loop. Upon defining $(Q,\circ)$ on $Q$ by \[ x\circ y = [(x\inv \ldiv y^2) x]^{1/2} = [(y^2)P_x]^{1/2}\,, \tag{$\circ$} \] we see that the bijection $(Q,\circ)\to (P_Q,\bullet)$; $x\mapsto P_x$ is an isomorphism of magmas. Thus $(Q,\circ)$ is a Bruck loop, the \emph{Bruck loop associated with the uniquely $2$-divisible automorphic loop $Q$}. We have established most of the following: \begin{proposition} \prplabel{bruckloop} Let $Q$ be a uniquely $2$-divisible automorphic loop. Then $(Q,\circ)$ defined by $(\circ)$ is a Bruck loop. Powers in $(Q,\circ)$ coincide with powers in $Q$. Any subloop of $Q$ which is closed under square roots is a subloop of $(Q,\circ)$. \end{proposition} \begin{proof} We already showed that $(Q,\circ)$ is a Bruck loop. Powers of $x$ in $(Q,\circ)$ correspond to powers of $P_x$ in $(P_Q,\bullet)$. But these coincide with powers of $P_x$ in $\Mlt(Q)$ \cite[Lem 4.5]{FKP}. By Corollary \corref{Ppowers}, we conclude that powers in $(Q,\circ)$ coincide with powers in $Q$. In Bruck loops, the left and right divisions can be expressed in terms of the multiplication and inversion: $x\ldiv_{\circ} y = x\inv \circ y$ and $x\rdiv_{\circ} y = y\inv\circ ((y\circ x)\circ y\inv)$. Thus the claim about subloops follows directly from ($\circ$). \end{proof} Note that $x\circ y = [(x\inv\ldiv y^2)x]^{1/2} = [x\inv\ldiv(y^2x)]^{1/2}$ by Proposition \prpref{commutes} \begin{proposition} Let $Q$ be a uniquely $2$-divisible automorphic loop. Then the core $(Q,\ast)$ is a quasigroup. \end{proposition} \begin{proof} This follows immediately from the unique $2$-divisibility, the fact that $(Q,\circ)$ is a loop, and the observation $x\ast y = (x\circ y^{-1/2})^2$. \end{proof} The \emph{left multiplication group} $\LMlt(Q)$ of a loop $Q$ is the group $\sbl{L_x\ |\ x\in Q}\le\Mlt(Q)$. \begin{lemma} \lemlabel{LMltBruck} Let $Q$ be a uniquely $2$-divisible automorphic loop with associated Bruck loop $(Q,\circ)$. Then $\LMlt(Q,\circ)$ is conjugate in $\Sym(Q)$ to $\sbl{P_Q}$. \end{lemma} \begin{proof} Let $\sigma : Q\to Q; x\mapsto x^2$ denote the squaring permutation. For each $x\in Q$, the left translation $y\mapsto x\circ y$ is just $\sigma P_x \sigma^{-1}$. This establishes the desired result. \end{proof} We will need the following easy observation later. \begin{lemma} \lemlabel{bruck-aut} Let $Q$ be a uniquely $2$-divisible automorphic loop with associated Bruck loop $(Q,\circ)$. Then $\Aut(Q)\leq \Aut(Q,\circ)$. In particular, every inner mapping of $Q$ acts as an automorphism of $(Q,\circ)$. \end{lemma} Next, we prove the Lagrange and Cauchy Theorems for automorphic loops of odd order. First, we must show that for finite automorphic loops, the notions of unique $2$-divisibility and having odd order coincide. In fact, this is true more generally for finite power-associative loops. \begin{lemma} \lemlabel{odd-2div} Let $Q$ be a finite loop with two-sided inverses. \begin{enumerate} \item[(i)] If $Q$ is uniquely $2$-divisible, then $Q$ has odd order. \item[(ii)] If $Q$ has odd order and the AAIP, then $Q$ has no elements of order $2$. If $Q$ is also power-associative, then $Q$ is uniquely $2$-divisible. \end{enumerate} \end{lemma} \begin{proof} Suppose $Q$ is uniquely $2$-divisible. Then the inversion mapping $J$ fixes only the identity element. Since $J$ has order $2$, the set of nonidentity elements of $Q$ must have even order, and so $Q$ has odd order. This proves (i). Now assume $Q$ has odd order and the AAIP, and suppose $c\in Q$ satisfies $c^2 = 1$. By the AAIP, if $xy = c$ then $c = c\inv = (xy)\inv = y\inv x\inv$. Thus the set $K = \setof{(x,y)}{xy = c}$ is invariant under the mapping $\phi: Q^2\to Q^2;(x,y)\mapsto (y\inv,x\inv)$. Since $\phi$ is involutive and $|K|$ is odd, $\phi$ has a fixed point $(x,y)\in K$. This point satisfies $x\inv = y$, so that $1 = x x\inv = c$. This establishes the first part of (ii), and the remaining assertion is clear. \end{proof} \begin{corollary} \corlabel{odd-2div} A finite automorphic loop is uniquely $2$-divisible if and only if it has odd order. \end{corollary} \begin{corollary} \corlabel{even} Let $Q$ be a finite automorphic loop of even order. Then $Q$ contains an element of order $2$. \end{corollary} \begin{proof} Otherwise, every element of $Q$ would have odd order, so that $Q$ would be uniquely $2$-divisible, and hence have odd order. \end{proof} \begin{lemma} \lemlabel{subloops} Let $Q$ be an automorphic loop of odd order with associated Bruck loop $(Q,\circ)$. If $S$ is a subloop of $Q$, then $S$ is a subloop of $(Q,\circ)$. \end{lemma} \begin{proof} In this case, the square root of any element is a positive integer power of that element, and so subloops are closed under taking square roots. Then Proposition \prpref{bruckloop} applies. \end{proof} \begin{theorem}[Lagrange Theorem] \thmlabel{lagrange} Let $Q$ be an automorphic loop of odd order. If $S\leq Q$, then $|S|$ divides $|Q|$. \end{theorem} \begin{proof} By Lemma \lemref{subloops}, $S$ is a subloop of the associated Bruck loop $(Q,\circ)$. The result follows from \cite[Cor. 4]{Gl1}. \end{proof} Note that Theorem \thmref{lagrange}, sometimes called the weak Lagrange property, implies what is known as the strong Lagrange property for automorphic loops of odd order: if $T\leq S\leq Q$, then $|T|$ divides $|S|$. This is because subloops of automorphic loops of odd order are themselves automorphic loops of odd order. \begin{theorem}[Cauchy Theorem] \thmlabel{cauchy} Let $Q$ be an automorphic loop of odd order. If a prime $p$ divides $|Q|$, then $Q$ contains an element of order $p$. \end{theorem} \begin{proof} By \cite[Coro 1, p. 394]{Gl1}, the associated Bruck loop $(Q,\circ)$ contains an element of order $p$ and thus so does $Q$ by Proposition \prpref{bruckloop}. \end{proof} \begin{corollary} \corlabel{orderpisgroup} Every automorphic loop of prime order is a group. \end{corollary} \begin{proof} This is trivial for $p = 2$, while for $p$ odd, it follows from Theorem \thmref{cauchy}. \end{proof} \section{A correspondence with Lie rings} \seclabel{correspondence} Following Wright \cite{Wright2}, if $(A,+,\cdot)$ is an algebra (over some field), define $(A,\diamond)$ by $x\diamond y = x + y - xy$. By \cite[Prop. 8]{Wright2}, $(A,\diamond)$ is a loop if and only if the mappings $y\mapsto y-yx$, $y\mapsto y-xy$ are bijections of $A$. We will now specialize this construction to Lie rings, and establish its partial inverse. Recall that a \emph{Lie ring} $(Q,+,[\cdot,\cdot])$ is an abelian group $(Q,+)$ such that the bracket $[\cdot,\cdot]$ is biadditive, satisfies the Jacobi identity $[x,[y,z]]+[y,[z,x]]+[z,[x,y]]=0$, and is alternating, that is, $[x,x]=0$. Consequently, Lie rings are skew-symmetric, $[x,y]=-[y,x]$. As usual, for $x\in Q$ define $\mathrm{ad}(x):Q\to Q$; $y\mapsto [y,x]$. Thanks to skew-symmetry, the mappings from Wright's construction take on the form \begin{align*} r_x &= \mathrm{id}_Q-\mathrm{ad}(x);\; y\mapsto y-[y,x],\\ \ell_x &= \mathrm{id}_Q+\mathrm{ad}(x);\; y\mapsto y+[y,x]. \end{align*} Note that all $r_x$, $\ell_x$ are homomorphisms of $(Q,+)$. In this context, Wright's construction can be stated as follows: \begin{lemma}\label{Lm:LieWright} Let $(Q,+,[\cdot,\cdot])$ be a Lie ring. Then $(Q,\diamond)$ defined by \begin{displaymath} x\diamond y = x + y - [x,y]\tag{$\diamond$} \end{displaymath} is a loop (with neutral element $0$) if and only if $(Q,+,[\cdot,\cdot])$ satisfies \eqref{Eq:Wright}, that is, if and only if the mappings $r_x$, $\ell_x$ are invertible for every $x\in Q$. For $x\in Q$ let $R_x^\diamond$, $L_x^\diamond$ be the right and left translation by $x$ in the groupoid $(Q,\diamond)$. Then \begin{align*} &y\diamond x = yR_x^\diamond = x+yr_x,\\ &x\diamond y = yL_x^\diamond = x+y\ell_x. \end{align*} If $(Q,\diamond)$ is a loop then also \begin{align*} y(R_x^\diamond)^{-1} &= (y-x)r_x^{-1},\\ y(L_x^\diamond)^{-1} &= (y-x)\ell_x^{-1},\\ R_y^\diamond R_z^\diamond (R_{y\diamond z}^\diamond)^{-1} &= r_yr_zr_{y\diamond z}^{-1},\\ L_y^\diamond L_z^\diamond (L_{z\diamond y}^\diamond)^{-1} &= \ell_y\ell_z\ell_{z\diamond y}^{-1},\\ R_y^\diamond (L_y^\diamond)^{-1} &= r_y\ell_y^{-1}. \end{align*} \end{lemma} \begin{proof} The first part of the statement is a special case of Wright's result, and the formulae for the translations $R_x^\diamond$, $L_x^\diamond$ follow from ($\diamond$) and skew-symmetry. For the rest of the proof suppose that $(Q,\diamond)$ is a loop. We immediately get the formulae for $(R_x^\diamond)^{-1}$ and $(L_x^\diamond)^{-1}$. Finally, $xR_y^\diamond R_z^\diamond (R_{y\diamond z}^\diamond)^{-1} = (z+(y+xr_y)r_z - (y\diamond z))r_{y\diamond z}^{-1} = (z+yr_z+xr_yr_z - (z+yr_z))r_{y\diamond z}^{-1} = xr_yr_zr_{y\diamond z}^{-1}$, $xL_y^\diamond L_z^\diamond (L_{z\diamond y}^\diamond)^{-1} = (z+(y+x\ell_y)\ell_z - (z\diamond y))\ell_{z\diamond y}^{-1} = (z+y\ell_z + x\ell_y\ell_z - (z+y\ell_z))\ell_{z\diamond y}^{-1} = x\ell_y\ell_z\ell_{z\diamond y}^{-1}$, and $xR_y^\diamond (L_y^\diamond)^{-1} = ((y+xr_y)-y)\ell_y^{-1} = xr_y\ell_y^{-1}$. \end{proof} The Lie ring construction sometimes yields automorphic loops: \begin{proposition} \label{Pr:LieToAut} Let $(Q,+,[\cdot,\cdot])$ be a Lie ring satisfying \eqref{Eq:Wright} and \eqref{Eq:WrightAutomorphic}. Then $(Q,\diamond)$ defined by $(\diamond)$ is an automorphic loop, and the commutant and nuclei of $(Q,\diamond)$ are given by \begin{align*} C(Q,\dia) &= \setof{a\in Q}{2[a,x]=0,\ \forall x\in Q}\\ N_{\lambda}(Q,\dia) &= \setof{a\in Q}{[[a,x],y] = 0,\ \forall x,\,y\in Q}\\ N_{\mu}(Q,\dia) &= \setof{a\in Q}{[[x,y],a] = 0,\ \forall x,\,y\in Q}. \end{align*} In particular, $(Q,\diamond)$ is a group if and only if $[[x,y],z]=0$ for all $x$, $y$, $z\in Q$. \end{proposition} \begin{proof} By Lemma \ref{Lm:LieWright}, $(Q,\diamond)$ is a loop. For all $x$, $y$, $z\in Q$, we have \begin{align*} x r_z\dia y r_z &= x - [x,z] + y - [y,z] - [x-[x,z],y-[y,z]] \\ &= x + y - [x,y] - [x + y,z] + [x,[y,z]] + [[x,z],y] - [[x,z],[y,z]] \\ &= x + y - [x,y] - [x + y,z] + [[x,y],z] \\ &= (x\dia y) r_z \,, \end{align*} where we have used both the Jacobi identity and the condition \eqref{Eq:WrightAutomorphic} in the third equality. Thus for each $z\in Q$ we have $r_z\in\Aut(Q,\dia)$. Similarly, $\ell_z\in\Aut(Q,\dia)$. By Lemma \eqref{Lm:LieWright}, the standard generators $R_y^\diamond R_z^\diamond (R_{y\diamond z}^\diamond)^{-1}$, $L_y^\diamond L_z^\diamond (L_{z\diamond y}^\diamond)^{-1}$ and $R_y^\diamond (L_y^\diamond)^{-1}$ of $\Inn(Q,\diamond)$ are elements of $\sbl{r_x,\,\ell_x\ |\ x\in Q}\le \Aut(Q,\dia)$, and hence $(Q,\diamond)$ is an automorphic loop. The characterization of the commutant is clear from ($\dia$). For the nuclei, we compute \[ \big((x\dia y)\dia z\big) - \big(x\dia (y\dia z)\big) = [[x,y],z] - [x,[y,z]] = [[x,z],y]\,, \] using the Jacobi identity. Thus a triple $x$, $y$, $z$ associates in $(Q,\dia)$ if and only if $[[x,z],y] = 0$. All remaining claims easily follow. \end{proof} \begin{corollary} \corlabel{glie} Let $Q$ be a Lie ring satisfying \eqref{Eq:Wright} and \eqref{Eq:WrightAutomorphic}, and let $(Q,\dia)$ be the corresponding automorphic loop. \begin{enumerate} \item[(i)] If $Q$ has characteristic $2$, then $(Q,\dia)$ is commutative. \item[(ii)] If the abelian group $(Q,+)$ is uniquely $2$-divisible, then $C(Q,\diamond) = Z(Q,\diamond)$ is equal to the center of the Lie ring $Q$. \end{enumerate} \end{corollary} For the rest of this section we will be concerned with the question of whether it is possible to invert the construction of Proposition \ref{Pr:LieToAut} to obtain a Lie ring satisfying \eqref{Eq:Wright} and \eqref{Eq:WrightAutomorphic} from an automorphic loop. We identify suitable subclasses of Lie rings and automorphic loops when this is indeed the case. For the rest of this section we will deal with uniquely $2$-divisible automorphic loops $Q$ for which the associated Bruck loop $(Q,\circ)$ is a group, hence an abelian group. For $x$ in such a loop $Q$, define the inner mapping \[ \phi_x = R_x P_{x^{1/2}}\inv. \tag{$\phi$} \] We will make heavy use of the fact that $\phi_x\in \Aut(Q)$, often without explicit reference. \begin{lemma} \lemlabel{op_corres} Let $Q$ be a uniquely $2$-divisible automorphic loop for which the associated Bruck loop $(Q,\circ)$ is an abelian group. For all $x$, $y\in Q$, the following identities hold: \begin{align} xy &= (x)\phi_y \circ y, \eqnlabel{op_corres} \\ x\circ (y\inv)\phi_x &= y\inv \circ (x)\phi_y. \eqnlabel{flip} \end{align} \end{lemma} \begin{proof} For all $x$, $y\in Q$, \[ x^2 y^2 = (x^2)R_{y^2} P_y\inv P_y = (x^2)\phi_{y^2} P_y = [((x)\phi_{y^2})^2]P_y = [(x)\phi_{y^2} \circ y]^2\,. \] Since $(Q,\circ)$ is an abelian group and powers in $(Q,\cdot)$, $(Q,\circ)$ coincide, we have $x^2 y^2 = ((x)\phi_{y^2})^2 \circ y^2 = (x^2)\phi_{y^2}\circ y^2$. Replacing $x$ with $x^{1/2}$ and $y$ with $y^{1/2}$, we obtain \eqnref{op_corres}. Now using AAIP, we have \[ (y\inv)\phi_{x\inv}\circ x\inv = y\inv x\inv = (xy)\inv = [(x)\phi_y\circ y]\inv = (x\inv)\phi_y \circ y\inv\,. \] Replacing $x$ with $x\inv$, we obtain \eqnref{flip}. \end{proof} \begin{lemma} \lemlabel{Pgroup} Let $Q$ be a uniquely $2$-divisible automorphic loop for which the associated Bruck loop $(Q,\circ)$ is an abelian group. Then $P_Q = \sbl{P_Q}$ is an abelian group isomorphic to $(Q,\circ)$. In particular, for all $x$, $y\in Q$, \begin{equation} \eqnlabel{Pgroup} P_x P_y = P_{x\circ y}\,. \end{equation} \end{lemma} \begin{proof} Since $(Q,\circ)$ is an abelian group, $(Q,\circ)\cong \LMlt(Q,\circ)\cong \sbl{P_Q}$ by Lemma \lemref{LMltBruck}. For \eqnref{Pgroup}, we have $P_{x\circ y} = P_x\bullet P_y = (P_x P_y^2 P_x)^{1/2} = (P_x^2 P_y^2)^{1/2} = P_x P_y$, since $\sbl{P_Q}$ is an abelian group. \end{proof} \begin{lemma} \lemlabel{phigenerate} Let $Q$ be a uniquely $2$-divisible automorphic loop for which the associated Bruck loop $(Q,\circ)$ is an abelian group. Then $\sbl{\phi_x\ |\ x\in Q} = \Inn(Q)$. \end{lemma} \begin{proof} One inclusion is obvious. We have \begin{equation} \eqnlabel{phigen1} R_{x,y} = R_x R_y R_{xy}\inv = \phi_x P_{x^{1/2}} \phi_y P_{y^{1/2}} P_{(xy)^{1/2}}\inv \phi_{xy}\inv = \phi_x \phi_y P_{(x^{1/2})\phi_y} P_{y^{1/2}} P_{(xy)^{1/2}}\inv \phi_{xy}\inv\,, \end{equation} since $\phi_y\in \Aut(Q)$. Now by \eqnref{Pgroup}, $P_{(x^{1/2})\phi_y} P_{y^{1/2}} = P_{(x^{1/2})\phi_y\circ y^{1/2}}$. By the fact that $(Q,\circ)$ is an abelian group and \eqnref{op_corres}, $(x^{1/2})\phi_y\circ y^{1/2} = [(x)\phi_y\circ y]^{1/2} = (xy)^{1/2}$. Thus \eqnref{phigen1} reduces to $R_{x,y} = \phi_x \phi_y \phi_{xy}\inv$. By \eqnref{RxyLxy}, $L_{x,y} = R_{x\inv,y\inv} = \phi_{x\inv} \phi_{y\inv} \phi_{x\inv y\inv}\inv$. Finally, \[ T_x = R_x L_x\inv = \phi_x P_{x^{1/2}} L_x\inv = \phi_x P_{x^{1/2}} P_{x\inv} R_{x\inv}\inv = \phi_x P_{x^{-1/2}} R_{x\inv}\inv = \phi_x \phi_{x\inv}\inv\,, \] where we used \eqnref{Pgroup} and $x^{1/2}\circ x\inv = x^{-1/2}$ in the fourth equality. It follows that $\Inn(Q)\leq \sbl{\phi_x\ |\ x\in Q}$. \end{proof} A Lie ring $(Q,+,[\cdot,\cdot])$ is said to be \emph{uniquely $2$-divisible} if the abelian group $(Q,+)$ is uniquely $2$-divisible. \begin{theorem}[Partial correspondence between Lie rings and automorphic loops] \thmlabel{correspondence} Suppose that $(Q,+,[\cdot,\cdot])$ is a uniquely $2$-divisible Lie ring satisfying \eqref{Eq:Wright} and \eqref{Eq:WrightAutomorphic}. Then $(Q,\dia)$ defined by \begin{displaymath} x\diamond y = x+y-[x,y] \end{displaymath} is a $2$-divisible automorphic loop whose associated Bruck loop $(Q,\circ)$ is an abelian group; in fact, $(Q,\circ) = (Q,+)$. Conversely, suppose that $(Q,\cdot)$ is a uniquely $2$-divisible automorphic loop whose associated Bruck loop $(Q,\circ)$ is an abelian group. Then $(Q,\circ,[\cdot,\cdot])$ defined by \[ [x,y] = x\circ y\circ (xy)\inv\tag{$[\cdot,\cdot]$} \] is a uniquely $2$-divisible Lie ring satisfying \eqref{Eq:Wright} and \eqref{Eq:WrightAutomorphic}. Furthermore, the two constructions are inverses of each other. Subrings (resp. ideals) of the Lie ring are subloops (resp. normal subloops) of the corresponding automorphic loop, and subloops (resp. normal subloops) closed under square roots are subrings (resp. ideals) of the corresponding Lie ring. \end{theorem} \begin{proof} Suppose that $(Q,+,[\cdot,\cdot])$ is a uniquely $2$-divisible Lie ring satisfying \eqref{Eq:Wright} and \eqref{Eq:WrightAutomorphic}. By Proposition \ref{Pr:LieToAut}, $(Q,\diamond)$ is an automorphic loop. Note that $x\diamond x = 2x$, $x^{-1}=-x$ and $x^{1/2} = \frac{1}{2}x$. The multiplication in the Bruck loop $(Q,\circ)$ associated with $(Q,\diamond)$ therefore has the form $x\circ y = ((2y)(L_{-x}^\diamond)^{-1}R_x^\diamond)\frac{1}{2} = ((2y)R_x^\diamond (L_{-x}^\diamond)^{-1})\frac{1}{2}$, where the second equality follows by Proposition \prpref{commutes}. Showing $x\circ y = x+y$ is therefore equivalent to proving $(2y)\diamond x = (2y)R_x^\dia = (2x+2y)L_{-x}^\dia = (-x)\dia(2x+2y)$. But $(2y)\dia x = 2y+x-[2y,x]=(-x)+(2x+2y)-[-x,2x+2y] = (-x)\dia (2x+2y)$. Conversely, suppose that $(Q,\cdot)$ is a uniquely $2$-divisible automorphic loop whose associated Bruck loop $(Q,\circ)$ is an abelian group. By \eqnref{op_corres}, we have $[x,y] = x\circ y\circ (xy)^{-1} = x\circ y \circ ((x)\phi_y\circ y)^{-1} = x\circ y \circ (x\inv)\phi_y \circ y\inv = x\circ (x\inv)\phi_y$. Since $(x\inv)\phi_x = x\inv$, we have $[x,x] = 1$. Next, \begin{displaymath} [x,y]\circ [y,x] = x\circ (x\inv)\phi_y \circ y \circ (y\inv)\phi_x = x\circ (y\inv)\phi_x \circ y\circ (x\inv)\phi_y = (x)\phi_y\circ y\inv \circ y\circ (x\inv)\phi_y = 1, \end{displaymath} where we have used \eqnref{flip} in the third equality and $\phi_y\in\Aut(Q)\le\Aut(Q,\circ)$ in the last equality. For biadditivity, we compute \begin{gather*} [x\circ y,z] = x\circ y \circ [(x\circ y)\inv]\phi_z = x\circ (x\inv)\phi_z \circ y\circ (y\inv)\phi_z = [x,z]\circ [y,z],\\ [x,y\circ z] = [y\circ z,x]\inv = ([y,x]\circ [z,x])\inv = [y,x]\inv\circ [z,x]\inv = [x,y]\circ [x,z]. \end{gather*} So far we have shown that $(Q,\circ,[\cdot,\cdot])$ is an alternating, biadditive (nonassociative) ring with underlying abelian group $(Q,\circ)$. In what follows the symbols $+$ and $-$ will refer to sums and differences of endomorphisms of $(Q,\circ)$. Rearranging the definition of $[\cdot,\cdot]$ and using the skew-symmetry, we have $xy = x\circ y\circ [x,y]\inv = y\circ (x)(\id_Q - \ad(y))$. Comparing this with \eqnref{op_corres}, we see that $\id_Q - \ad(x) = \phi_x$ and also $\id_Q + \ad(x) = \phi_{x\inv}$. In particular, property \eqref{Eq:Wright} holds. Now using biadditivity, we have \[ [(x)(\id_Q + \ad(z)),(y)(\id_Q + \ad(z))] = [x,y]\circ [x,[y,z]]\circ [[x,z],y] \circ [[x,z],[y,z]]\,, \] and also \[ [x,y](\id_Q + \ad(z)) = [x,y] \circ [[x,y],z]. \] Since $\id_Q + \ad(x)=\phi_x \in \Aut(Q)\leq \Aut(Q,[\cdot,\cdot])$, the results of these two calculations are equal. Canceling common terms and rearranging using skew-symmetry, we obtain \begin{equation} \eqnlabel{lietmp} [[x,y],z]\circ [[y,z],x]\circ [[z,x],y] = [[x,z],[y,z]] \end{equation} for all $x$, $y$, $z\in Q$. Since the left side of \eqnref{lietmp} is invariant under cyclic permutations of $x$, $y$, $z$, so is the right side, and so we have \begin{equation} \eqnlabel{lietmp2} [[x,z],[y,z]] = [[y,x],[z,x]] \end{equation} for all $x$, $y$, $z\in Q$. Replace $x$ in this last identity with $x\circ u$ and use biadditivity to get \[ [[x,z],[y,z]]\circ [[u,z],[y,z]] = [[y,x],[z,x]]\circ [[y,u],[z,x]] \circ [[y,x],[z,u]]\circ [[y,u],[z,u]]. \] Canceling terms on both sides using \eqnref{lietmp2}, we obtain $1 = [[y,u],[z,x]]\circ [[y,x],[z,u]]$ for all $x$, $y$, $z$, $u\in Q$. Taking $u = y$, we get $1 = [[y,x],[z,y]]$, which is equivalent to \eqref{Eq:WrightAutomorphic}. It follows that the right side of \eqnref{lietmp} is equal to $1$, and so the Jacobi identity holds. Therefore, $(Q,\circ,[\cdot,\cdot])$ is a Lie ring satisfying \eqref{Eq:Wright} and \eqref{Eq:WrightAutomorphic}. Let us now show that the two constructions are inverse to each other. Suppose that the constructions yield $(Q,\cdot)\mapsto (Q,\circ,[\cdot,\cdot])\mapsto (Q,\dia)$. Then $x\dia y = x\circ y\circ[x,y]^{-1}$, and since $(Q,\circ)$ is an abelian group and $x\circ y\circ (xy)^{-1}=[x,y]$, we conclude that $x\dia y = xy$. In the other direction, let $(Q,+,[\cdot,\cdot])\mapsto (Q,\diamond)\mapsto (Q,\circ,\lceil \cdot,\cdot\rceil)$, where $(Q,\circ)$ is the Bruck loop associated with $(Q,\dia)$. We have already shown that $(Q,\circ)=(Q,+)$. Then $\lceil x,y\rceil = x\circ y\circ (x\dia y)^{-1} = x+y-(x+y-[x,y])=[x,y]$. Finally, we show the correspondence of substructures. Suppose that $(Q,\cdot)$ corresponds to $(Q,+,[\cdot,\cdot])$. Lemma \ref{Lm:LieWright} shows that the three loop operations of $(Q,\cdot)$ (in fact, of $(Q,\dia)$, but $(Q,\dia)=(Q,\cdot)$ here) can be expressed in terms of $+$ and $[\cdot,\cdot]$. If $S$ is a subring of $(Q,+,[\cdot,\cdot])$, then since $S$ is closed under $+$ and $[\cdot,\cdot]$, it is a subloop of $(Q,\cdot)$. If $S$ is an ideal of $(Q,+,[\cdot,\cdot])$, then it is invariant under the mappings $\id_Q - \ad(x) = \phi_x$ for all $x\in Q$ and hence $S$ is invariant under $\Inn(Q,\cdot)$ by Lemma \lemref{phigenerate}. If $S$ is a subloop of $(Q,\cdot)$ closed under square roots, then by Proposition \prpref{bruckloop} $S$ is a subgroup of $(Q,\circ)$. Therefore $S$ is a subring of $(Q,+,[\cdot,\cdot])$ by definition of the bracket. Finally, if $S$ is a normal subloop of $(Q,\cdot)$, then $S$ is invariant under all mappings $\id_Q - \ad(x) = \phi_x$. But then $(S)\ad(x)\subseteq S$ for all $x\in Q$, and so $S$ is an ideal of $(Q,+,[\cdot,\cdot])$. \end{proof} We conclude with the observation that in the uniquely $2$-divisible case the condition \eqref{Eq:WrightAutomorphic} already implies that $(Q,+,[\cdot,\cdot])$ is solvable of derived length at most $2$. \begin{lemma}\label{Lm:SolvabilityFollows} Let $(Q,+,[\cdot,\cdot])$ be a uniquely $2$-divisible Lie ring. Then $Q$ satisfies \eqref{Eq:WrightAutomorphic} if and only if $[[Q,Q],[Q,Q]]=0$. \end{lemma} \begin{proof} Clearly, if $[[Q,Q],[Q,Q]]=0$ then \eqref{Eq:WrightAutomorphic} follows. For the converse, suppose that $[[x,y],[z,y]]=0$ for all $x$, $y$, $z\in Q$. Replacing $y$ with $y+u$ and then using \eqref{Eq:WrightAutomorphic} itself to cancel terms, we obtain $[[x,y],[z,u]]+[[x,u],[z,y]] = 0$ or by skew-symmetry, \begin{equation} \eqnlabel{lietmp4} [[x,y],[z,u]] = [[x,u],[y,z]] \end{equation} for all $x$, $y$, $z$, $u\in Q$. Now if we apply the identity \eqnref{lietmp4} to its own right hand side, we obtain $[[x,u],[y,z]] = [[x,z],[u,y]]$, which together with \eqnref{lietmp4} gives \begin{equation} \eqnlabel{lietmp5} [[x,y],[z,u]] = [[x,z],[u,y]] \end{equation} for all $x$, $y$, $z$, $u\in Q$. On the other hand, \eqnref{lietmp4} is equivalent to \begin{equation} \eqnlabel{lietmp6} [[x,y],[z,u]] = [[u,x],[z,y]] \end{equation} for all $x$, $y$, $z$, $u\in Q$, using skew-symmetry. If we apply \eqnref{lietmp6} to its own right hand side, we obtain $[[u,x],[z,y]] = [[y,u],[z,x]] = - [[x,z],[u,y]]$, using skew-symmetry in the last equality. This together with \eqnref{lietmp6} gives \begin{equation} \eqnlabel{lietmp7} [[x,y],[z,u]] = -[[x,z],[u,y]] \end{equation} for all $x$, $y$, $z$, $u\in Q$. Comparing \eqnref{lietmp5} and \eqnref{lietmp7}, we have \[ 2[[x,y],[z,u]] = 0 \] for all $x$, $y$, $z$, $u\in Q$. Since $(Q,+)$ is uniquely $2$-divisible, it follows that $[[x,y],[z,u]] = 0$ for all $x$, $y$, $z$, $u\in Q$. \end{proof} \section{Nilpotency and Solvability} \seclabel{solvability} In this section we prove the Odd Order Theorem for automorphic loops together with two other corollaries of Theorem \thmref{correspondence}. We start with automorphic loops of prime power order. Let $p$ be a prime. By Corollary \corref{orderpisgroup}, an automorphic loop of order $p$ is isomorphic to $\mathbb Z_p$. The following result was first obtained by Cs\"org\H{o} \cite{Csorgo1}, using her signature method of connected transversals. We can now give a short proof based on Theorem \thmref{correspondence}. A proof that is both short and elementary remains elusive. \begin{theorem}(Cs\"org\H{o}) \thmlabel{orderp2} Let $p$ be a prime. Every automorphic loop of order $p^2$ is a group. \end{theorem} \begin{proof} Let $Q$ be an automorphic loop of order $p^2$. Every loop of order $4$ is associative \cite{Pflugfelder}, so assume $p>2$. Bruck loops of order $p^2$ are groups \cite{Burn}. If $(Q,\circ)$ is cyclic, then so is $Q$, so assume $(Q,\circ)$ is elementary abelian. Theorem \thmref{correspondence} and Lemma \ref{Lm:SolvabilityFollows} give an associated solvable Lie ring $(Q,\circ,[\cdot,\cdot])$ of derived length at most $2$. Since $(Q,\circ)$ is an elementary abelian, $(Q,\circ,[\cdot,\cdot])$ is a $2$-dimensional Lie algebra over $GF(p)$. Over any field, there are, up to isomorphism, only two $2$-dimensional Lie algebras, one abelian and the other nonabelian \cite{Humph}. The nonabelian Lie algebra of dimension $2$ has a basis $\{x,y\}$ such that $[x,y] = y$. But then $y(\id + \ad(x)) = 0$ so that condition \eqref{Eq:Wright} is not satisfied. Thus $(Q,\circ,[\cdot,\cdot])$ must be an abelian Lie algebra, that is, $[x,y] = 0$ for all $x$, $y\in Q$. Then $xy = x\circ y$, that is, $Q$ is an abelian group. \end{proof} Commutative automorphic loops of order $p^k$ are centrally nilpotent when $p$ is an odd prime \cite{Csorgo1,JKV3}. Commutative automorphic loops of order $p^3$ were classified up to isomorphism in \cite{dBGV}. There are additional nonassociative noncommutative automorphic loops of order $p^3$, $p$ and odd prime. A class of such loops with trivial nucleus was obtained in \cite{JKV3}. In particular, when $p$ is an odd prime, automorphic loops of order $p^3$ need not be centrally nilpotent. Here we present the construction of \cite{JKV3} in a new way, using the corresponding Lie algebras: \begin{example} \exmlabel{orderp3} Let $F$ be a field and fix $A\in GL(2,F)$. On $Q = F\times F^2$, define an operation $[\cdot,\cdot]$ by \[ [(a,x),(b,y)] = (0,(ay-bx)A) \] for all $a$, $b\in F$, $x$, $y\in F^2$. (Note that we think of elements of $F^2$ as row vectors so that $A$ acts on the right.) Then it is straightforward to verify that $(Q,+,[\cdot,\cdot])$ is a Lie algebra satisfying \eqref{Eq:WrightAutomorphic}. Let $r_x = \mathrm{id}_Q - \mathrm{ad}(x)$, $\ell_x = \mathrm{id}_Q+\mathrm{ad}(x)$ be as before. In block matrix form, we have \begin{align*} (a,x)r_{(b,y)} &= (a,x+(bx-ay)A) = (a,x)\begin{pmatrix} 1 & -yA \\ 0 & I + bA \end{pmatrix} \\ \intertext{and} (b,y)\ell_{(a,x)} &= (b,y+(bx-ay)A) = (b,y)\begin{pmatrix} 1 & xA \\ 0 & I - aA \end{pmatrix}\,, \end{align*} where $I$ is the $2\times 2$ identity matrix. Thus condition \eqref{Eq:Wright} will hold precisely when $\det(I + \mu A)\neq 0$ for all $\mu\in F$, that is, when the characteristic polynomial of $A$ has no roots in $F$. (See \cite{JKV3} for an interpretation of this in terms of anisotropic planes.) Assume this property now holds for $A$. We show that the left/right nucleus of the corresponding loop $(Q,\dia)$ is trivial. By Proposition \ref{Pr:LieToAut}, $N_{\lambda}(Q,\dia)$ consists of all elements $(a,x)$ such that $[[(a,x),(b,y)],(c,z)] = (0,c(ay-bx)A) = (0,0)$ for all $b$, $c\in F$, $y$, $z\in F^2$. Thus $c(ay-bx)A = 0$. Since $A\in GL(2,F)$ and taking $c\neq 0$, we have $ay=bx$ for all $b\in F$, $y\in F^2$. Taking $b\neq 0$, $y=0$ implies $x = 0$, while taking $b=0$, $y\neq 0$ implies $a=0$. Thus $N_{\lambda}(Q,\dia)$ is trivial. Consider the particular case $F = GF(p)$. If $p = 2$, then by Corollary \corref{glie}, we obtain a commutative automorphic loop $(Q,\dia)$ of exponent $2$ and order $8$. There is precisely one such loop with trivial center, first constructed in \cite{JKV2}. As discussed in \cite{JKV3}, if $p = 3$, then this construction gives two isomorphism classes of (noncommutative) automorphic loops depending on the choice of $A$, while if $p = 5$, there are three isomorphism classes. For $p > 5$, it is conjectured that there are precisely three isomorphism classes \cite[Conj. 6.5]{JKV3}. \end{example} Returning to general automorphic loops of order $p^3$, $p$ odd prime, there is much that is still unknown, but we can at least say that for $p = 3$, such automorphic loops are necessarily given by the construction of Proposition \ref{Pr:LieToAut}: \begin{lemma} \lemlabel{order27} Let $Q$ be an automorphic loop of order $27$ and exponent $3$. Then $Q$ is constructed from a Lie algebra satisfying \eqref{Eq:Wright} and \eqref{Eq:WrightAutomorphic} by the construction $(\dia)$. \end{lemma} \begin{proof} Every Bruck loop of exponent $3$ is a commutative Moufang loop \cite{Rob}. Moufang loops of order $3^n$ for $n\leq 3$ are associative. Thus the associated Bruck loop $(Q,\circ)$ is an elementary abelian $3$-group. By Theorem \thmref{correspondence}, we have an associated solvable Lie ring $(Q,\circ,[\cdot,\cdot])$ satisfying \eqref{Eq:Wright}, \eqref{Eq:WrightAutomorphic}. Since $(Q,\circ)$ is elementary abelian, $(Q,\circ,[\cdot,\cdot])$ is a Lie algebra over $GF(3)$. By Theorem \thmref{correspondence}, $(Q,\cdot)$ is equal to the loop $(Q,\dia)$ obtained from $(Q,\circ,[\cdot,\cdot])$ by $(\dia)$. \end{proof} Lemma \lemref{order27} cannot be easily extended to Bruck loops of order $p^3$ and exponent $p$ for $p > 3$ because there are nonassociative Bruck loops of such orders. We now start working toward the Odd Order Theorem. If $Q$ is a loop and $S\leq Q$, the \emph{relative multiplication group of} $S$, denoted by $\Mlt(Q;S)$, is the subgroup of $\Mlt(Q)$ generated by all $R_x$, $L_x$, $x\in S$. The \emph{relative inner mapping group of} $S$ is $\Inn(Q;S) = (\Mlt(Q;S))_1 = \Mlt(Q;S)\cap \Inn(Q)$. \begin{lemma} \lemlabel{odd-subloop} Let $Q$ be a finite automorphic loop of odd order. A subloop $S$ of the associated Bruck loop $(Q,\circ)$ is a subloop of $Q$ if and only if $S h = S$ for every $h\in \Inn(Q;S)$. \end{lemma} \begin{proof} The ``only if'' direction is trivial, so assume the hypothesis of the converse assertion. Fix $u$, $v \in S$. Since powers agree in $(Q,\circ)$ and $Q$, we have $u\inv, v\inv\in S$. Set $w = v^{1/2}$ and note that $v\in S$ as well. By Lemma \lemref{bruck-aut}, $\Aut(Q,\cdot)\le\Aut(Q,\circ)$. Thus $S$ also contains \[ (u \circ w)^2 T_u = (uT_u \circ wT_u)^2 = (u \circ wT_u)^2 = (u\inv \ldiv [w T_u]^2) u = v T_u L_{u\inv}^{-1} R_u = v R_u^2 L_u\inv L_{u\inv}\inv , \] using \eqnref{RLLR}. Since $L_{u\inv} L_u \in \Inn(Q)$, $S$ also contains $v R_u^2 = (u\circ w)^2 T_u L_{u\inv} L_u$. By induction, $v R_u^{2k}\in S$ for all integers $k$. Now let $2n+1$ be the order of $u$. Then $R_u^{2n+1} \in \Inn(Q)$, and so $S$ contains $v R_u^{2n+1} R_u^{-2n} = v u$, and also $v R_u^{2n+1} R_u^{2(-n-1)} = v \rdiv u$. Thus $S$ is closed under multiplication and right division. By the AAIP, $S$ is also closed under left division, and hence is a subloop. \end{proof} \begin{lemma} \lemlabel{Bruck_characteristic} Let $Q$ be a uniquely $2$-divisible automorphic loop, and let $(Q,\circ)$ be the associated Bruck loop. Then every characteristic subloop of $(Q,\circ)$ is a normal subloop of $Q$. \end{lemma} \begin{proof} If $S$ is a characteristic subloop of $(Q,\circ)$, then by Lemma \lemref{bruck-aut}, $S$ is invariant under $\Inn(Q)$. By Lemma \lemref{odd-subloop}, $S$ is subloop of $Q$. \end{proof} \begin{theorem}[Odd Order Theorem] \thmlabel{ft} Every automorphic loop of odd order is solvable. \end{theorem} \begin{proof} Let $Q$ be a minimal counterexample. If $1 < S\propnormal Q$, then by minimality, both $S$ and $Q/S$ are solvable automorphic loops of odd order. This contradicts the nonsolvability of $Q$. Therefore $Q$ is simple. Let $(Q,\circ)$ be the associated Bruck loop and let $D$ denote the derived subloop of $(Q,\circ)$. By \cite[Thm. 14(b)]{Gl2}, $(Q,\circ)$ is solvable and so $D$ is a proper subloop. Since $D\Char(Q,\circ)$, it follows from Lemma \lemref{Bruck_characteristic} that $D\unlhd Q$. Since $Q$ is simple, $D = \{1\}$. Therefore $(Q,\circ)$ is an abelian group. Now let $p$ be a prime divisor of $|Q|$ and let $M_p=\{x\in Q\ |\ x^p=1\}$. Then $M_p\Char(Q,\circ)$, and so by Lemma \lemref{Bruck_characteristic} again, $M_p\unlhd Q$. By Theorem \thmref{cauchy}, $M_p$ is nontrivial, and so since $Q$ is simple, $M_p = Q$. Thus $Q$ has exponent $p$, $(Q,\circ)$ has exponent $p$ by Proposition \prpref{bruckloop}, and $(Q,\circ)$ is an elementary abelian $p$-group. By Theorem \thmref{correspondence}, $(Q,\circ,[\cdot,\cdot])$ defined by ($[\cdot,\cdot]$) is a Lie ring satisfying \eqref{Eq:Wright} and \eqref{Eq:WrightAutomorphic}. By Lemma \ref{Lm:SolvabilityFollows}, $(Q,\circ,[\cdot,\cdot])$ is solvable. Since $(Q,\circ)$ is an elementary abelian $p$-group, we may view $(Q,\circ,[\cdot,\cdot])$ as a finite dimensional Lie algebra over $GF(p)$. Since $Q$ is simple as a loop, Theorem \thmref{correspondence} also implies that $(Q,\circ,[\cdot,\cdot])$ is either simple as a Lie algebra or else is abelian. The former case contradicts the solvability of $(Q,\circ,[\cdot,\cdot])$, and so $(Q,\circ,[\cdot,\cdot])$ is abelian. But then $xy = x\circ y\circ [x,y] = x\circ y$, so that $Q$ is an abelian group, a contradiction with nonsolvability of $Q$. \end{proof} We remark that the proof of \cite[Thm. 14(b)]{Gl2} depends on the Feit-Thompson Odd Order Theorem for groups, and hence so does our proof of Theorem \thmref{ft}. \section{Finite Simple Automorphic Loops} \seclabel{simplicity} The main open problem in the theory of automorphic loops is the existence or nonexistence of a nonassociative finite simple automorphic loop, \emph{cf.}, Problem \prbref{simple}. By Theorem \thmref{ft} and by the main results of \cite{GKN}, such a loop would be noncommutative and of even order, though not a $2$-loop. Simple loops can be studied via primitive permutation groups thanks to this classic theorem of Albert \cite{albert}: \begin{proposition} \prplabel{albert} A loop $Q$ is simple if and only if $\Mlt(Q)$ is primitive. \end{proposition} \begin{lemma} \lemlabel{centralizer} Let $Q$ be a simple nonassociative automorphic loop with inversion map $J$. If $J\ne\mathrm{id}_Q$ then $C_{\Mlt(Q)}(J) = \Inn(Q)$. \end{lemma} \begin{proof} Since $Q$ is automorphic, $J$ commutes with every inner mapping. Therefore $\Inn(Q)\leq C_{\Mlt(Q)}(J)$. Since $\Mlt(Q)$ is primitive by Proposition \prpref{albert}, $\Inn(Q)$ is a maximal subgroup of $\Mlt(Q)$. Since $J \neq \id_Q$, there is $x\in Q$ such that $x\ne x^{-1}$, and so $xJL_x =1\ne x^{-2}=xL_xJ$. Hence $C_{Mlt(Q)}(J) \neq \Mlt{Q}$, and so the desired equality holds. \end{proof} Recall that the \emph{socle} $\Soc(G)$ of a group $G$ is the subgroup generated by the minimal normal subgroups of $G$. By the O'Nan-Scott Theorem \cite[Thm. 4.1A]{DM}, the analysis of a finite primitive group $G$ divides into two cases depending on whether or not $\Soc(G)$ is regular. \begin{proposition} Let $Q$ be a finite simple nonassociative automorphic loop. Then the socle $\Soc(\Mlt(Q))$ is not regular. \end{proposition} \begin{proof} Suppose $S = \Soc(\Mlt(Q))$ is regular. Recall that $J$ normalizes $\Mlt(Q)$ in $\Sym(Q)$ by Corollary \corref{AAIP}. Thus since $S$ is characteristic in $\Mlt(Q)$, $S$ is normalized by $J$. By Theorem \thmref{ft}, $|S| = |Q|$ is even. Thus $J$ fixes a nonidentity element $s \in S$. If $J \neq \mathrm{id}_Q$, then by Lemma \lemref{centralizer}, $s \in \Inn(Q)$. But then $(1)s = 1$, which contradicts the regularity of $S$. Therefore $J = \mathrm{id}_Q$ and so $Q$ has exponent $2$. By \cite[Thm. 6.2]{JKV1}, $Q$ has order a power of $2$ and then by \cite[Thm. 3]{GKN}, $Q$ is solvable, a contradiction. \end{proof} By the O'Nan-Scott Theorem, it follows that $\Mlt(Q)$ is of almost simple type, of diagonal type or of product type \cite{DM}. Although the classification of finite simple automorphic loops remains open, results from group theory about characteristic subgroups hold analogously for characteristic subloops of automorphic loops with essentially the same proofs (\emph{cf}. the closing remarks of \cite{GKN}). Part (ii) of the following result is \cite[Thm. 2.2(ii)]{BP}. \begin{theorem} \thmlabel{characteristic} Let $Q$ be an automorphic loop. \begin{enumerate} \item[(i)] If $T \Char S\normal Q$, then $T \normal Q$. \item[(ii)] Every characteristic subloop of $Q$ is normal. \item[(iii)] If $Q$ is finite and characteristically simple, then $Q$ is a direct product of isomorphic simple loops. \end{enumerate} \end{theorem} \begin{proof} Every inner mapping leaves $S$ invariant, hence acts as an automorphism of $S$. Since $T$ is characteristic in $S$, $T\varphi = T$ for all $\varphi\in \Inn(Q)$. This establishes (i), and (ii) follows from (i) by taking $S = Q$. Now suppose $Q$ is finite and characteristically simple, and let $S=S_1$ be a minimal normal subloop. Consider the orbit $\{S_1,\ldots,S_m\}$ of $S$ under $\Aut(Q)$. Each $S_i$, being the image of a minimal normal subloop of $Q$ under an automorphism, is also a minimal normal subloop of $Q$. Since each $S_i\cap S_j$ is normal in $Q$, it follows from minimality that the subloops $S_i$ intersect pairwise trivially. Thus $S_1\cdots S_m$ is a direct product \cite{Bruck}. Since automorphisms map the direct factors of $S_1\cdots S_m$ to each other, the direct product is characteristic in $Q$. Thus $Q=S_1\cdots S_m$ because $Q$ is characteristically simple. Since $S$ is both a minimal normal subloop and a direct factor of $Q$, $S$ must be simple. This establishes (iii). \end{proof} \begin{corollary} \corlabel{minimal_normal} A minimal normal subloop of a finite automorphic loop is a direct product of isomorphic simple loops. \end{corollary} \begin{proof} If $S$ is a minimal normal subloop of $Q$, then by Theorem \thmref{characteristic}(i), $S$ is characteristically simple, and we are done by Theorem \thmref{characteristic}(iii). \end{proof} \begin{proposition} Let $Q$ be an automorphic loop. \begin{enumerate} \item[(i)] $Q^{(n)} \normal Q$ for each $n > 0$. \item[(ii)] If $Q$ is solvable, then the \emph{derived series} $Q \backnormal Q' \backnormal Q'' \backnormal \cdots \backnormal Q^{(n)} = 1$ is a normal series, that is, $Q^{(k)} \normal Q$ for all $k > 0$. \end{enumerate} \end{proposition} \begin{proof} Each $Q^{(k)}$ is characteristic in $Q^{(k-1)}$ for all $k\geq 1$. By Theorem \thmref{characteristic}(i), each $Q^{(n)} \normal Q$. This proves (i), and (ii) follows from (i). \end{proof} \section{Split Middle Nuclear Extensions} \seclabel{mid_nuc} In this brief section we will examine automorphic loops which are split extensions by their middle nuclei. The following proposition shows that this notion can be defined in either of the usual group theoretic ways. \begin{proposition} \prplabel{split} Let $Q$ be a loop with normal middle nucleus $N_{\mu}=N_{\mu}(Q)$. The following conditions are equivalent. \begin{enumerate} \item[(i)] The natural homomorphism $\eta: Q\to Q/N_{\mu}$ splits, that is, there is a homomorphism $\sigma: Q/N_{\mu} \to Q$ such that $\sigma\eta = \id_{Q/N_{\mu}}$. \item[(ii)] There exists a subloop $S$ of $Q$ such that $Q = SN_{\mu}$ and $S\cap N_{\mu} = 1$. \end{enumerate} \end{proposition} \begin{proof} Assume (i) holds. Let $S = \sigma(Q/N_{\mu})$, which is a subloop of $Q$ since $\sigma$ is a homomorphism. Clearly $S\cap N_{\mu} = 1$. For $x\in Q$, let $s = (x)\eta\sigma = (x N_{\mu})\sigma$ and let $a = s\backslash x$. Then $(a)\eta = N_{\mu}$, that is, $a\in \ker(\eta) = N_{\mu}$. Therefore $Q = SN_{\mu}$ and (ii) holds. Assume (ii) holds. Suppose $sa = tb$ for $s,t\in S$ and $a$, $b\in N_{\mu}$. Then $s = tb\cdot a\inv = t\cdot ba\inv$ since $a$, $b\in N_{\mu}$. Hence $t\ldiv s = ba\inv \in S\cap N_{\mu} = 1$, and so $t=s$ and $b=a$. Thus each $x\in Q$ has a unique factorization $x = sa$ for some $s\in S$, $a\in N_{\mu}$. In particular, the subloop $S$ is a complete set of left coset representatives of $N_{\mu}$. Therefore setting $(sN_{\mu})\sigma = s$ for each $s\in S$ yields a well-defined map $\sigma : Q/N_{\mu} \to Q$ with $(sN_{\mu})\eta\sigma = sN_{\mu}$. Finally $\sigma$ is a homomorphism by the definition of coset multiplication. \end{proof} We will say that an automorphic loop $Q$ is a \emph{split middle nuclear extension} (of $S$ by $N_{\mu}$) if either, and hence both, of the conditions of Proposition \prpref{split} hold. In automorphic loops, the multiplication in a split middle nuclear extension has a very specific form: \begin{proposition} \prplabel{mid_nuc} Let $Q$ be an automorphic loop. For all $a$, $b\in N_{\mu}(Q)$ and all $x$, $y\in Q$, \begin{equation} \eqnlabel{mid_nuc_mult} xa\cdot yb = xy\cdot ((a)T_y\cdot b)L_{y,x}\,. \end{equation} \end{proposition} \begin{proof} First we prove \begin{equation} \eqnlabel{mid_nuc_tmp} a\cdot yb = ay\cdot b\,. \end{equation} Since $b\in N_{\mu}=N_{\mu}(Q)$, we have $T_b = R_b L_{b\inv}$. Thus we compute \begin{displaymath} ay\cdot b = b\cdot (ay)T_b = b[(a)T_b\cdot (y)T_b] = b[b\inv ab\cdot b\inv yb] = b[b\inv a\cdot yb] = (b\cdot b\inv a)\cdot yb = a\cdot yb, \end{displaymath} where we have used $T_b\in \Aut(Q)$ in the second equality, $b,b\inv\in N_{\mu}$ in the fourth equality and $b\inv a\in N_{\mu}$ in the fifth equality. This establishes \eqnref{mid_nuc_tmp}. For \eqnref{mid_nuc_mult}, we compute \[ xa\cdot yb = x(a\cdot yb) \byeqn{mid_nuc_tmp} x(ay\cdot b) = x\cdot (y\cdot (a)T_y)b = x\cdot y((a)T_y\cdot b) = xy\cdot ((a)T_y\cdot b)L_{y,x}\,, \] where we have used $(a)T_y\in N_{\mu}$ (since $N_{\mu} \normal Q$ by Proposition \prpref{nuclei}) in the fourth equality. \end{proof} \begin{corollary} \corlabel{split_mult} Let $Q$ be an automorphic loop which is a split middle nuclear extension $Q = SN_{\mu}$. Then for all $s$, $t\in S$, $a$, $b\in N_{\mu}$, \begin{equation} \eqnlabel{split_mult} sa\cdot tb = st\cdot ((a)T_t\cdot b)L_{t,s}\,, \end{equation} where the right hand side is the unique factorization of the left side into an element $st\in S$ and an element $((a)T_t\cdot b)L_{t,s}\in N_{\mu}$. \end{corollary} Just as split extensions of groups (internal semidirect products) lead naturally to external semidirect products, so do split middle nuclear extensions of automorphic loops lead to an ``external'' construction of automorphic loops. The input data are a loop $S$, a group $N$, a mapping $\phi : S\to \Aut(N)$ satisfying $(1)\phi=1$ and a mapping $\alpha : S\times S\to \Aut(N)$ satisfying $(1,s)\alpha = (s,1)\alpha = 1$ for all $s\in S$. On $Q := S\times N$, we define operations by \begin{align*} (s,a)\cdot (t,b) &= (st,(a^{(t)\phi}b)^{(t,s)\alpha})\,, \\ (s,a)\ldiv (t,b) &= (s\ldiv t, (a\inv)^{(s\ldiv t)\phi} b^{((s\ldiv t,s)\alpha)\inv})\,, \\ (s,a)\rdiv (t,b) &= (s\rdiv t, (a^{((t,s\rdiv t)\alpha)\inv}b\inv)^{((t)\phi)\inv})\,. \end{align*} Then it is easy to show $(Q,\cdot,\ldiv,\rdiv)$ is a loop with neutral element $(1,1)$. To get an automorphic loop, it is necessary that $S$ be automorphic and there are various conditions which must be satisfied by $\phi$ and $\alpha$. It is straightforward to find these conditions by simply calculating inner mappings in $Q$ and assuming them to be automorphisms. However, the calculations and the conditions themselves are both lengthy and unenlightening in their full generality. Since we are only going to examine a special case in detail in the next section, we omit the general construction. \section{Dihedral Automorphic Loops} \seclabel{dihedral} We begin with a construction of automorphic loops motivated by Corollary \corref{split_mult}. \begin{proposition} Let $(A,+)$ be an abelian group and fix $\alpha\in \Aut(A)$. Let $\Dih(A,\alpha)$ be defined on $\mathbb Z_2\times A$ by \begin{equation} \eqnlabel{gen_dihedral} (i,u)\cdot (j,v) = (i + j, ((-1)^ju + v)\alpha^{ij}). \end{equation} Then $(\Dih(A,\alpha),\cdot)$ is an automorphic loop. If $\alpha\neq\id_A$, then $N_{\mu} = \{0\}\times A \cong A$. \end{proposition} \begin{proof} Throughout the proof, the exponent of $\alpha$ in \eqnref{gen_dihedral} is calculated in $\mathbb Z_2$. Clearly $(0,0)$ is the neutral element. Setting \begin{align*} (i,u)\ldiv (j,v) &= (i+j, v\alpha^{-i(j+i)} - (-1)^{i+j}u ),\\ (i,u)\rdiv (j,v) &= (i+j, (-1)^j (u \alpha^{-(i+j)j} - v), \end{align*} it is straightforward to show that $\ldiv$ and $\rdiv$ satisfy the properties of divisions in a loop. The generalized conjugation $T_{(i,u)}$ is given by \[ (j,v)T_{(i,u)} = (j, (-1)^i v + (1-(-1)^j)u )\,, \] as can be readily checked. Note that this is independent of $\alpha$. We check that this is an automorphism. First, \begin{align*} [(j,v)\cdot(k,w)]T_{(i,u)} &= (j + k, ((-1)^k v + w)\alpha^{jk} )T_{(i,u)}\\ &= (j+k, (-1)^i((-1)^k v + w)\alpha^{jk} + (1-(-1)^{j+k})u)\\ &= (j+k, (-1)^{i+k}v\alpha^{jk} + (-1)^iw\alpha^{jk} + (1-(-1)^{j+k})u). \end{align*} On the other hand, $(j,v)T_{(i,u)}\cdot (k,w)T_{(i,u)} =$ \begin{align*} &= (j, (-1)^i v + (1-(-1)^j)u )\cdot (k, (-1)^i w + (1-(-1)^k)u )\\ &= (j+k, [(-1)^k((-1)^i v + (1-(-1)^j)u) + (-1)^i w + (1-(-1)^k)u]\alpha^{jk} )\\ &= (j+k, (-1)^{i+k}v\alpha^{jk} + (-1)^iw\alpha^{jk} + h), \end{align*} where \[ h = [(-1)^k(1-(-1)^j)+(1-(-1)^k)]u\alpha^{jk} = (1-(-1)^{j+k})u\alpha^{jk}\,. \] Checking all four possibilities, we see that $(1 - (-1)^{j+k})u\alpha^{jk} = (1 - (-1)^{j+k})u$ for $j,k\in \mathbb{Z}_2$. Thus $T_{(i,u)}$ is an automorphism. Next, we check that the left inner mappings $L_{(j,v),(i,u)}$ are automorphisms. A lengthy calculation gives \[ (k,w)L_{(j,v),(i,u)} = (k,[(-1)^{j+k}u(\alpha^{-jk}-\id_A) + w]\alpha^{ij})\,. \] Note that this is independent of $v$. We have \begin{align*} (k,w)&L_{(j,v),(i,u)}\cdot (\ell,x)L_{(j,v),(i,u)} \\ &= (k,[(-1)^{j+k}u(\alpha^{-jk}-\id_A) + w]\alpha^{ij})\cdot (\ell,[(-1)^{j+\ell}u(\alpha^{-j\ell}-\id_A) + x]\alpha^{ij})\\ &=(k+\ell,\{(-1)^{\ell}[(-1)^{j+k}u(\alpha^{-jk}-\id_A) + w] + (-1)^{j+\ell}u(\alpha^{-j\ell}-\id_A) + x \}\alpha^{ij}\alpha^{k\ell})\\ &= (k+\ell, [(-1)^{\ell}w + x + q]\alpha^{ij}\alpha^{k\ell} ), \end{align*} where \begin{align*} q &= (-1)^{\ell}u[(-1)^{j+k}(\alpha^{-jk}-\id_A) + (-1)^j(\alpha^{-j\ell}-\id_A)]\\ &= (-1)^{j + k + \ell}u(\alpha^{-jk}-\id_A + (-1)^k(\alpha^{-j\ell}-\id_A ))\\ &= (-1)^{j + k + \ell}u(\alpha^{-j(k+\ell)} - \id_A). \end{align*} The last equality follows by checking all possible values of $j$, $k$, $\ell\in \mathbb{Z}_2$. On the other hand, we compute \begin{align*} [(k,w)&\cdot (\ell,x)]L_{(j,v),(i,u)} = \\ &= (k+\ell,[(-1)^{\ell}w + x]\alpha^{k\ell})L_{(j,v),(i,u)}\\ &= (k+\ell,\{(-1)^{j+k+\ell}u(\alpha^{-j(k+\ell)}-\id_A) + [(-1)^{\ell}w + x]\alpha^{k\ell}\}\alpha^{ij})\\ &= (k+\ell,\{(-1)^{j+k+\ell}u(\alpha^{-j(k+\ell)}-\id_A)\alpha^{-k\ell} +(-1)^{\ell}w + x\}\alpha^{k\ell}\alpha^{ij}). \end{align*} Now observe that $u(\alpha^{-j(k+\ell)}-\id_A)\alpha^{-k\ell} = u(\alpha^{-j(k+\ell)}-\id_A)$ for all $j$, $k$, $\ell\in \mathbb{Z}_2$ just by checking all possibilities. Thus we see that $L_{(j,v),(i,u)}$ is an automorphism. Applying Proposition \prpref{halfcheck}, we have shown that $\Dih(A,\alpha)$ is an automorphic loop. It remains to characterize the middle nucleus when $\alpha\neq\id_A$. We have that $(j,v)\in N_m$ if and only if $(k,w) = (k,w)L_{(j,v),(i,u)}$ for all $i$, $k\in \mathbb{Z}_2$, $u$, $w\in A$. Thus matching second components, we require \begin{equation} \eqnlabel{mid_nuc_chk} [(-1)^{j+k}u(\alpha^{-jk} - \id_A) + w]\alpha^{ij} = w \end{equation} for all $i$, $k\in \mathbb{Z}_2$, $u$, $w\in A$. Taking $u = 0$, $i = 1$, we must have $w\alpha^j = w$ for all $w\in A$. Thus $\alpha^j = \id_A$. Since $\alpha\neq\id_A$, we must have $j=0$. On the other hand, since \eqnref{mid_nuc_chk} is independent of $v$, it is clear that $(0,v)\in N_{\mu}$. This completes the proof. \end{proof} We call the loops $\Dih(A,\alpha)$ \emph{generalized dihedral automorphic loops}. $\Dih(A,\id_A)$ is the usual generalized dihedral group determined by the abelian group $A$. If $A = \mathbb{Z}$, then $\Aut(A) = \mathbb{Z}^* = \{\pm 1\}$. In this case we write $D_{\infty}(c) = \Dih(\mathbb{Z},c)$ where $c = \pm 1$ and refer to these loops as \emph{infinite dihedral automorphic loops}. If $A = \mathbb{Z}_n$, then $\Aut(A) = \mathbb{Z}_n^*$, the group of integers in $\{1,\ldots,n-1\}$ coprime to $m$. We write $D_{2n}(c) = \Dih(\mathbb{Z}_n,c)$ where $c\in \mathbb{Z}_n^*$ and refer to these loops simply as \emph{dihedral automorphic loops}. In $D_{\infty}(c)$ or $D_{2n}(c)$, the multiplication specializes as follows: \begin{equation}\label{Eq:Dihedral} (i,j)\cdot (k,\ell) = (i + k, c^{ik}((-1)^k j + \ell)), \end{equation} where $c\in \{\pm 1\}$ in the former case and $c\in \mathbb{Z}_n^*$ in the latter case. We now show that different values of the parameter $c$ give nonisomorphic dihedral automorphic loops, and we calculate their automorphism groups. \begin{lemma}\label{Lm:DihedralGens} Let $Q=D_{2n}(c)$. Then \begin{enumerate} \item[(i)] $(0,1)^m = (0,m)$ for every $m\in \mathbb Z$, and $\mathbb Z_n\cong 0\times \mathbb Z_n\le Q$, \item[(ii)] $|(1,x)|=2$ for every $x\in\mathbb Z_n$, \item[(iii)] $(1,0)\cdot(0,y) = (1,y)$ for every $y\in\mathbb Z_n$, and $Q = \langle (0,1),\,(1,0)\rangle$. \end{enumerate} \end{lemma} \begin{proof} (i) Since automorphic loops are power-associative, the power $(0,1)^m$ is well-defined for every $m\in\mathbb Z$. The claim holds for $m=0$, since $(0,0)$ is the neutral element of $Q$. Suppose the claim holds for some $m\ge 0$. Then $(0,1)^{m+1} = (0,1)^m\cdot(0,1) = (0,m)\cdot(0,1) = (0,m+1)$. Since $(0,-m)\cdot(0,m)=(0,0)$, it follows that $(0,1)^{-m} = (0,-m)$. The rest is clear. (ii) For any $x\in\mathbb Z_n$ we have $(1,x)\cdot(1,x) = (0,c(-x+x)) = (0,0)$. (iii) The formula $(1,0)\cdot(0,y) = (1,y)$ follows immediately from \eqref{Eq:Dihedral}. Then $Q=\langle (0,1),(1,0)\rangle$ follows from (i). \end{proof} By Lemma \ref{Lm:DihedralGens}, a loop homomorphism $f:D_{2n}(c)\to Q$ is determined by its values $(0,1)f$, $(1,0)f$. If $n>2$ then $0\times\mathbb Z_n$ is the unique subloop of $D_{2n}(c)$ isomorphic to $\mathbb Z_n$, by Lemma \ref{Lm:DihedralGens}(iii). Hence, if $f:D_{2n}(c)\to D_{2n}(d)$ is an isomorphism, it follows that $(0,1)f=(0,\alpha)$ for some $\alpha\in \mathbb Z_n^*$, and $(1,0)f=(1,\beta)$ for some $\beta\in\mathbb Z_n$. Using Lemma \ref{Lm:DihedralGens} again, we then have \begin{align*} (0,x)f &= ((0,1)^x)f = ((0,1)f)^x = (0,\alpha)^x = (0,x\alpha),\\ (1,x)f &= ((1,0)\cdot(0,x))f = (1,0)f\cdot(0,x)f = (1,\beta)\cdot(0,x\alpha) = (1,\beta+x\alpha) \end{align*} for every $x\in\mathbb Z_n$. Given any $\alpha\in\mathbb Z_n^*$, $\beta\in\mathbb Z_n$, let us denote the mapping $f:D_{2n}(c)\to D_{2n}(d)$ satisfying $(0,x)f = (0,x\alpha)$, $(1,x)f = (1,\beta+x\alpha)$ for all $x\in\mathbb Z_n$ by $f_{\alpha,\beta}$. (Note that the definition of $f_{\alpha,\beta}$ does not require knowledge of $c$, $d$, so we will consider $f_{\alpha,\beta}$ to be a mapping from $D_{2n}(c)$ to $D_{2n}(d)$ for any $c$, $d\in\mathbb Z_n^*$.) \begin{lemma}\label{Lm:NearHomomorphism} Let $c$, $d\in\mathbb Z_n^*$, $\alpha\in Z_n^*$ and $\beta\in \mathbb Z_n$. Then $f=f_{\alpha,\beta}:D_{2n}(c)\to D_{2n}(d)$ is a bijection that satisfies $((0,x)\cdot(0,y))f = (0,x)f\cdot(0,y)f$, $((0,x)\cdot(1,y))f = (0,x)f\cdot(1,y)f$ and $((1,x)\cdot(0,y))f = (1,x)f\cdot(0,y)f$ for every $x$, $y\in\mathbb Z_n$. Moreover, $f$ is an isomorphism if and only if $c=d$. \end{lemma} \begin{proof} Since $\alpha\in\mathbb Z_n^*$, it is clear from the definition of $f=f_{\alpha,\beta}$ that it is a bijection $D_{2n}(c)\to D_{2n}(d)$. For $x$, $y\in\mathbb Z_n$ we have $((0,x)\cdot(0,y))f = (0,x+y)f = (0,(x+y)\alpha) = (0,x\alpha)\cdot(0,y\alpha) = (0,x)f\cdot(0,y)f$, $((0,x)\cdot(1,y))f = (1,-x+y)f = (1,\beta + (-x+y)\alpha) = (0,x\alpha)\cdot(1,\beta+y\alpha) = (0,x)f\cdot(1,y)f$, and $((1,x)\cdot(0,y))f = (1,x+y)f = (1,\beta + (x+y)\alpha) = (1,\beta+x\alpha)\cdot(0,y\alpha) = (1,x)f\cdot(0,y)f$. Finally, we have $((1,x)\cdot(1,y))f = (0,c(-x+y))f = (0,c(-x+y)\alpha)$, while $(1,x)f\cdot (1,y)f = (1,\beta + x\alpha)\cdot(1,\beta+y\alpha) = (0,d(-(\beta+x\alpha)+\beta+y\alpha)) = (0,d(-x+y)\alpha)$, so $f$ is an isomorphism if and only if $c=d$. \end{proof} \begin{corollary}\label{Cr:DihedralNonisomorphic} For an integer $n\ge 2$, the loops $D_{2n}(c)$, $c\in\mathbb Z_n^*$ are pairwise nonisomorphic. \end{corollary} \begin{proposition} \label{Pr:AutomorphismGroup} Let $c\in\mathbb Z_n^*$ and $Q=D_{2n}(c)$. Then $\mathrm{Aut}(Q)$ is isomorphic to the holomorph $\mathrm{Aut}(\mathbb Z_n)\rtimes \mathbb Z_n = \mathbb Z_n^*\rtimes \mathbb Z_n$ with multiplication $(\alpha,\beta)(\gamma,\delta) = (\alpha\gamma,\beta + \alpha\delta)$. \end{proposition} \begin{proof} By the discussion preceding Lemma \ref{Lm:NearHomomorphism}, every automorphism of $Q$ is of the form $f_{\alpha,\beta}$ for some $\alpha\in\mathbb Z_n^*$, $\beta\in\mathbb Z_n$. By Lemma \ref{Lm:NearHomomorphism}, every such mapping $f_{\alpha,\beta}$ is an automorphism of $Q$. Now, if $\gamma\in\mathbb Z_n^*$, $\delta\in\mathbb Z_n$ and $x\in\mathbb Z_n$, we have $(0,x)f_{\gamma,\delta}f_{\alpha,\beta} = (0,x\gamma)f_{\alpha,\beta} = (0,x\gamma\alpha) = (0,x\alpha\gamma) = (0,x)f_{\alpha\gamma,\beta+\alpha\delta}$ and $(1,x)f_{\gamma,\delta}f_{\alpha,\beta} = (1,\delta+x\gamma)f_{\alpha,\beta} = (1,\beta+(\delta +x\gamma)\alpha) = (1,\beta+\alpha\delta + x\alpha\gamma) = (1,x)f_{\alpha\gamma,\beta+\alpha\delta}$. \end{proof} Results analogous to \ref{Lm:DihedralGens}--\ref{Pr:AutomorphismGroup} hold for the infinite dihedral automorphic loops $D_\infty(c)$, with every occurrence of $\mathbb Z_n$ replaced with $\mathbb Z$, and $2n$ replaced with $\infty$. Commutative automorphic loops with middle nuclei of index $2$ were studied in detail in \cite{JKV2}. In the next result we examine the noncommutative case under the assumption that the middle nucleus is cyclic. \begin{proposition} \prplabel{index2} Let $Q$ be a noncommutative automorphic loop with cyclic middle nucleus $N_{\mu}(Q) = \langle b\rangle$, and suppose that $Q$ is a split middle nuclear extension $Q = \langle a\rangle \langle b\rangle$ where $a^2 = 1$. If $Q$ is infinite, then $Q \cong D_{\infty}(c)$ for some $c\in \{\pm 1\}$. If $Q$ is finite, then $Q \cong D_{2n}(c)$ for some $n\in \mathbb{N}$ and some $c\in \mathbb{Z}_n^*$. \end{proposition} \begin{proof} Since $T_a^2 = \id_Q$ by Lemma \lemref{RxyLxy} (see \eqnref{Tinv}), we must have $(b)T_a = b$ or $(b)T_a = b\inv$ by the normality of $\langle b\rangle$ in $Q$. If the former situation holds, then $(a^i b^j)T_a = a^i b^j$ for all $i=0$, $1$ and all $j$ since $T_a$ is an automorphism. Therefore $T_a$ fixes every point of $Q$ and hence $a\in C(Q)$. It follows that $(a^i b^j)T_b = a^i b^j$ for all $i=0$, $1$ and all $j$ since $T_b$ is an automorphism. Thus $b\in C(Q)$. Therefore $C(Q) = Q$, that is, $Q$ is commutative, a contradiction. It follows that $(b)T_a = b\inv$. We have that $T_1 = L_{1,1} = L_{1,a} = L_{a,1} = \id_Q$, and so referring to \eqnref{split_mult}, we see that the multiplication in $Q$ is entirely determined by the automorphism $L_{a,a} \upharpoonright \langle b\rangle$. If $\langle b\rangle \cong \mathbb{Z}$, then $\Aut(\langle b\rangle) \cong \mathbb{Z}^* = \{\pm 1\}$ and there are two possible values for $L_{a,a} \upharpoonright \langle b\rangle$ determined by $(b)L_{a,a} = b^c$ where $c = \pm 1$. If $\langle b\rangle \cong \mathbb{Z}_n$, then $\Aut(\langle b\rangle) \cong \mathbb{Z}_n^*$, and the possible values for $L_{a,a} \upharpoonright \langle b\rangle$ are given by $(b)L_{a,a} = b^{c}$ where $c\in \mathbb{Z}_n^*$. In either case, we thus have $(b)L_{a^i,a^k} = b^{c^{ik}}$ for $i, k = 0,1$. Fixing $c\in \mathbb{Z}^*$ or $\mathbb{Z}_n^*$, it follows from the preceding discussion that \eqnref{split_mult} specializes to the present setting as follows: \begin{equation} \eqnlabel{index2} a^i b^j \cdot a^k a^{\ell} = a^{i+k} b^{c^{ij}((-1)^k j + \ell)}\,, \end{equation} for all $i$, $k\in \mathbb{Z}_2$, $j$, $\ell\in \mathbb{Z}$ or $\mathbb{Z}_n$. Finally, for $m = \infty$ or $2n$, define $\psi : D_m(c)\to Q$ by $(i,j)\psi = a^i b^j$. It is straightforward to check that $\psi$ is an isomorphism using \eqref{Eq:Dihedral} and \eqnref{index2}. \end{proof} As an application, we have the following classification results. \begin{theorem} \thmlabel{order2modd} Let $Q$ be a finite automorphic loop with a cyclic subgroup of odd order $n$ and of index $2$. Then either $Q$ is a cyclic group or $Q\cong D_{2n}(c)$ for some $c\in \mathbb{Z}_n^*$. \end{theorem} \begin{proof} Let $\langle b\rangle$ be a cyclic subgroup of order $n$. This subloop is normal in $Q$ since it has index $2$. By Corollary \corref{even}, $Q$ also has an element $a$ of order $2$. By \eqnref{Tinv}, $(b)T_a = b$ or $(b)T_a = b\inv$, and by the same argument as in the proof of Proposition \prpref{index2}, we see that the former case leads to $Q$ being commutative. If $Q$ is commutative, then by \cite[Thm. 5.1]{JKV1}, $Q$ is isomorphic to the direct product $\mathbb{Z}_2 \times \mathbb{Z}_n\cong \mathbb{Z}_{2n}$. Thus we assume from now on that $Q$ is noncommutative, and so $(b)T_a = b\inv$. It remains to show that $\langle b\rangle$ is the middle nucleus of $Q$. Since $Q$ is the disjoint union of $\langle b\rangle$ and $a\langle b\rangle$, every element of $Q$ has a unique representation in the form $a^i b^j$, $i = 0,1$, $0\leq j < n$. Thus to show $\langle b\rangle \subseteq N_\mu(Q)$, we must show $(a^i b^j\cdot b^k)\cdot a^{\ell} b^r = a^i b^j\cdot (b^k\cdot a^{\ell} b^r)$ for all $0\leq i$, $\ell \leq 1$, $0\leq j$, $k$, $r < n$. Our first step is to prove \begin{equation} \eqnlabel{basecase} bab = a\,. \end{equation} Set $c = b^{(n+1)/2}$ so that $c^2 = b$. We use \eqnref{PRnorm} to get $(x\inv)P_{xy} = (x\inv)R_{x\inv}\inv P_y L_x = xy^2$ for all $x$, $y\in Q$. Take $x = a\rdiv c$ and $y = c$ in this to get $(a\rdiv c)b = (a\rdiv c)c^2 = [(a/c)\inv ]P_a = [(a/c)\inv ]T_a$ because $P_a = T_a$ since $a^2 = 1$. We record this as $(a\rdiv c)b = [(a\rdiv c)\inv ]T_a$, and use this identity twice in the following: \begin{align*} b\cdot(a\rdiv c)b &= b\cdot [(a/c)\inv ]T_a = (b\inv)T_a \cdot [(a/c)\inv ]T_a = [b\inv (a\rdiv c)\inv]T_a\\ &= [((a\rdiv c)b)\inv]T_a = [((a\rdiv c)b)T_a]\inv = (((a\rdiv c)\inv)T_a T_a)^{-1} = a\rdiv c, \end{align*} where we also used $T_a\in \Aut(Q)$ in the third and fifth equalities, and AAIP in the fourth. Hence $aR_c = aR_c R_b L_b = aR_b L_b R_c$ by Proposition \prpref{commutes}. Canceling, we obtain \eqnref{basecase}. Recall that we work under the assumption $(b)T_a= b^{-1}$. By \eqnref{basecase}, we have $a = bab = (a\cdot (b)T_a)b = ab\inv\cdot b$. Thus the automorphism $L_a R_b L_a\inv R_b\inv$ fixes $b\inv$ and hence fixes each $b^k$, that is, $ab^k\cdot b = ab^{k+1}$ for $0\leq k < n$. Then the automorphism $L_{b^k} L_a L_{ab^k}\inv$ fixes $b$ and hence fixes each $b^r$, that is, $ab^k\cdot b^r = ab^{k+r}$ for $0\leq k, r < n$. Since $ab^k \cdot a = a\cdot b^ka$ (by Proposition \prpref{commutes}), $L_{b^k} L_a L_{ab^k}\inv$ also fixes $a$ and hence fixes each $a^{\ell}b^r$, that is, $ab^k\cdot a^{\ell}b^r = a(b^k\cdot a^{\ell}b^r)$ for $0\le \ell \le 1$, $0 \leq k,r < n$. On the other hand, by \eqnref{basecase} again, we have $a = bab = b((b)T_a\inv\cdot a) = b((b)T_a\cdot a) = b\cdot b\inv a$. Dualizing the arguments of the preceding paragraph, we get $b^j (b^k\cdot a^{\ell}b^r) = b^{j+k}\cdot a^{\ell}b^r$ for $\ell = 0,1$, $0\leq j,k,r < n$. Combining this with the preceding paragraph, we see that $R_{b^k} R_{a^{\ell}b^r} R_{b^k\cdot a^{\ell}b^r}\inv$ fixes both $a$ and each $b^j$. It follows that $(a^i b^j\cdot b^k)\cdot a^{\ell} b^r = a^i b^j\cdot (b^k\cdot a^{\ell} b^r)$ for $0\leq i,\ell\leq 1$, $0\leq j,k,r < n$, as desired. We have shown that $\langle b\rangle \subseteq N_\mu(Q)$. If $ab^i\in N_\mu(Q)$ for any $i$, then $a\in N_\mu(Q)$ since $N_\mu(Q)$ is a subloop. But then $Q = N_\mu(Q)$, a contradiction. Therefore $\langle b\rangle = N_\mu(Q)$. By Proposition \prpref{index2}, we have the desired result. \end{proof} Recently, P. Cs\"{o}rg\H{o} was able to establish the following result by group-theoretic means: \begin{theorem}[Elementwise Lagrange Theorem \cite{Csorgo2}] \thmlabel{elementwise_Lagrange} Let $Q$ be a finite automorphic loop and let $a\in Q$. Then the order of $a$ divides the order of $Q$. \end{theorem} \begin{corollary} [Automorphic loops of order $2p$] \corlabel{order2p} Let $Q$ be an automorphic loop of order $2p$ where $p$ is an odd prime. Then $Q\cong D_{2p}(c)$ for some integer $1 \le c < p$, or $Q\cong \mathbb Z_{2p}$. Thus there are precisely $p$ automorphic loops of order $2p$, including the cyclic group $\mathbb Z_{2p}$ and the dihedral group $D_{2p}$. \end{corollary} \begin{proof} By Corollary \corref{even}, $Q$ has an element $a$ of order $2$. If every element of $Q$ had order $2$, then by \cite[Thm. 8]{GKN}, $Q$ itself would have order a power of $2$, a contradiction. By Theorem \thmref{elementwise_Lagrange}, every element of $Q$ has order dividing $2p$. Thus $Q$ must have an element $b$ of order $p$. Since $\langle a\rangle \cap \langle b\rangle = 1$, the desired isomorphism now follows from Theorem \thmref{order2modd}. The $p-1$ dihedral automorphic loops $D_{2p}(c)$, $c\in\mathbb Z_p^*$ are pairwise nonisomorphic by Corollary \ref{Cr:DihedralNonisomorphic}. \end{proof} \section{Open Problems} \seclabel{problems} The main open problem in the theory of automorphic loops is the following: \begin{problem} \prblabel{simple} Does there exist a (finite) simple, nonassociative automorphic loop? \end{problem} Also open are the Lagrange, Cauchy, Sylow and Hall theorems. \begin{problem} \prblabel{lagrange} Let $Q$ be a finite automorphic loop and let $S\leq Q$. Does $|S|$ divide $|Q|$? \end{problem} \begin{problem} \prblabel{sylow} Let $Q$ be a finite automorphic loop. \begin{enumerate} \item[(i)] For each prime $p$ dividing $|Q|$, does $Q$ have an element of order $p$? \item[(i)] For each prime $p$ dividing $|Q|$, does $Q$ have a Sylow $p$-subloop? \item[(i)] If $Q$ is solvable and if $\pi$ is a set of primes, does $Q$ have a Hall $\pi$-subloop? \end{enumerate} \end{problem} \begin{acknowledgment} Our investigations were aided by the automated deduction tool \textsc{Prover9} and the finite model builder \textsc{Mace4}, both developed by McCune \cite{McCune}, and the LOOPS package \cite{LOOPS} for GAP \cite{GAP}. We thank Ian Wanless for the idea behind the proof of Lemma \lemref{odd-2div}(ii). We thank G\'{a}bor Nagy for remarking that standard group theory facts about characteristic subgroups should hold for characteristic subloops of automorphic loops. \end{acknowledgment}
\section{Appendixes}
\section{Introduction} The nonlinear interplay of gravity, fluid dynamics, and radiative cooling lies at the heart of the continuing challenge to predict the properties of the first astrophysical objects of the universe. Scale separation, which is one of the most important prerequisites for robust analytic or numerical calculations, does not apply: whereas the kinetic energy budget is dominated by large-scale modes, fragmentation and cooling instabilities grow fastest on those scales with the strongest density fluctuations, i.e., those close to the Jeans scale. In order to capture the relevant degrees of freedom, grid-based hydrodynamical simulations employ adaptive mesh refinement over many orders-of-magnitude. In addition to the masses and multiplicity of the first stars, the question of whether and when supermassive black holes (SMBH) can form by direct collapse of a primordial gas cloud belongs to the key topics of current research \citep{2002ApJ...569..558O,2003ApJ...596...34B,2006ApJ...652..902S,2006MNRAS.370..289B,2008MNRAS.391.1961D,2008arXiv0803.2862D,2010MNRAS.402.1249S,2010ApJ...712L..69S,2011MNRAS.411.1659L}. Massive primordial haloes assembled at redshifts 10-15 are the plausible sites to host the first galaxies formed at the end of cosmic dark ages. They are also potential candidates for the formation of intermediate mass black holes through direct collapse \citep{1984ARA&A..22..471R, 2003ApJ...596...34B, 2004MNRAS.354..292K, 2006MNRAS.370..289B, 2009ApJ...702L...5B, 2010A&ARv..18..279V}. Thus, their understanding is a matter of great astrophysical interest. The first stars, on the other hand, are expected to form in so-called minihalos with $\rm 10^5-10^6$~M$_\odot$ \citep{2002Sci...295...93A, 2004PASP..116..103B, Yoshida08}. In a recent study, \citet{2012ApJ...745..154T} found that an increased resolution per Jeans length results in non-convergence of global properties. They discovered that enhanced Jeans resolution produces higher infall velocities, increased temperatures as well as decreased content of molecular hydrogen. As convergence has not been achieved even at the highest resolutions, major uncertainties are present concerning the expected accretion rates, temperature distributions and the fragmentation behavior. It is therefore important to assess whether similar restrictions apply to other systems, in particular the so-called atomic cooling haloes. In the study presented here, we therefore assess the convergence behavior employing numerical simulations with a resolution of $\rm 16$, $\rm 32$ and $\rm 64$ cells per Jeans length. We focus on halos exposed to strong Lyman Werner radiation, as these are the candidates for black hole formation via direct collapse. We note that these are the highest resolution studies to date, with previous studies typically employing a resolution of $\rm 16$ cells per Jeans length \citep{2008ApJ...682..745W, 2009MNRAS.393..858R, 2010MNRAS.402.1249S}. The need for high-resolution investigations has been previously derived in different contexts. In particular, \citet{2011ApJ...731...62F} reported that the turbulent energy as in collapsing gas clouds converges only for a resolution of at least $\rm 32$ cells per Jeans length. The turbulent energy in these clouds is released from the gravitational potential, due to the initial deviations from spherical symmetry \citep{1953ApJ...118..513H,1982ApJ...258L..29S,2010A&A...520A..17K,2010ApJ...712..294E}. It is well known from the field of contemporary star formation that turbulence has important implications for the density PDF, clump statistics, and angular momentum transport in gravitationally unstable gas clouds (e.g., \citep{1981MNRAS.194..809L,2004RvMP...76..125M,2007ARA&A..45..565M,2012arXiv1209.2856F}). During high-redshift structure formation, turbulence was shown to play a major role in so-called minihalos. In particular, it regulates the angular momentum transport and delays the formation of a disk \citep{2002Sci...295...93A, 2004PASP..116..103B, Yoshida08}, but also influences the fragmentation behavior \citep{Clark11, Smith11, Greif12}. In more massive halos, the presence of turbulence was reported by \cite{2008MNRAS.387.1021G} and \cite{2008ApJ...682..745W}. While the implications were not explored in detail, it is expected that it can influence the formation of intermediate mass black holes through its impact on disk formation and the angular momentum distribution. We note that the presence of disks is a central assumption in direct collapse models \citep{2004MNRAS.354..292K, 2006MNRAS.370..289B,2007NCimR..30..293L,2009ApJ...702L...5B}, and an efficient means of angular momentum transport is generally required to allow the formation of a massive central object. \citet{2009ApJ...702L...5B} therefore invoked the presence of nested bar-like instabilities to provide sufficient means for angular momentum transport. While our simulations do not exactly support this generation mechanism, they show that turbulence is efficiently produced during gravitational collapse. Such turbulence will not only influence angular momentum transport, but also amplify existing weak magnetic fields via the small scale dynamo \citep{1968JETP...26.1031K,1998MNRAS.294..718S,2005PhR...417....1B,Schobera,2010A&A...522A.115S,2010ApJ...721L.134S,2011ApJ...731...62F, 2012ApJ...745..154T,Schoberb}. The enhanced magnetic field strength could exert an additional pressure, and further contribute to the angular momentum transport. More importantly, this initial growth via the small-scale dynamo provides a strong initial field on which the $\rm \alpha-\Omega$ dynamo can act to produce the coherent fields observed today \citep{Beck96, Arshakian09}. We note that an accurate modeling of turbulence not only requires high numerical resolution, but also a consistent treatment of the unresolved scales. For this purpose, we explore the implications of a turbulence subgrid-scale (SGS) model for the turbulent kinetic energy \citep{2009ApJ...707...40M,2011MNRAS.414.2297I,2011A&A...528A.106S}. Its main features are non-ideal terms in the fluid equations produced by an effective turbulent viscosity, an effective turbulent pressure, and fully time and space dependent dissipation of turbulent kinetic energy into thermal energy. All of these terms make order unity contributions for transonic flows, which is the case here. In this study, we therefore employ high-resolution studies of massive halos to explore their central properties. A central question is in particular for which properties convergence can be achieved, which is directly relevant for the direct collapse scenario concerning the formation of intermediate mass black holes. Our paper is organized as follows. In the next section, we describe the simulation setup and the numerical methods employed. In the 3rd section of the paper, we present the results obtained in this study. In the last section of this article, we summarize our main results and confer our conclusions. \section{Numerical Methods and simulation details} A modified version of the Enzo code including the subgrid-scale model for unresolved turbulent fluctuations (see below for a description) has been used to perform the simulations presented in this work. Enzo is an adaptive mesh refinement (AMR), parallel, grid-based cosmological hydrodynamics code \citep{2004astro.ph..3044O,2007arXiv0705.1556N}. It can run on massively parallel systems and has been used for a wide variety of astrophysical applications. The message passing interface (MPI) is used to achieve portability and scalability on different systems. The computational domain is discretized into nested grid cells. It has two exchangeable grids, a uniform grid and a block-structured adaptive grid. We use a split hydro solver with a 3rd order piece-wise parabolic (PPM) method for hydrodynamical calculations. The dark matter N-body dynamics is solved using the particle-mesh technique. A multigrid Poisson solver is employed for the self-gravity computations. \subsection{Initial conditions} The simulations are started with the cosmological initial conditions generated from Gaussian random fields. We employ the inits package available with the public version of the Enzo code to create nested grid initial conditions. Our simulations start at redshift $\rm z=99$ with a top grid resolution of $\rm 128^{3}$ cells and we select the massive halo at redshift 15 using the halo finder of \cite{2011ApJS..192....9T}. Two initial nested levels of refinement are subsequently added each with a resolution of $\rm 128^{3}$ cells. Our simulation box has a cosmological size of 1 Mpc $\rm h^{-1}$ and is centered on the massive halo. In total, we initialize 6291456 particles to compute the evolution of the dark matter dynamics and have a final dark matter resolution of 300 $\rm M_{\Theta}$. While our dark matter halo is thus well-resolved, we note that additional fluctuations could be present in case of a higher resolution in dark matter. The parameters for creating the initial conditions and the distribution of baryonic and dark matter components are taken from the WMAP seven years data \citep{2011ApJS..192...14J}. We further allow additional 27 levels of refinement in the central 62 kpc region of the halo during the course of simulation. It gives us a total effective resolution of 3 AU in comoving units. The resolution criteria used in these simulations are based on the Jeans length, the gas over-density and the particle mass resolution. The grid cells matching these requirement are marked for a refinement. The simulations conducted in this work mandated the Jeans length resolution of 16, 32 and 64 cells throughout their evolution. This criterion was applied during the course of simulations to ensure that all physical processes like shock waves and Truelove criterion \citep{1997ApJ...489L.179T} are well resolved. We stop the simulations after they reach the maximum refinement level and start to violate the Jeans criterion. The results at later stages would not be reliable. \subsection{Chemistry} To include the primordial non-equilibrium chemistry, the rate equations of the following 9 species: $\rm H,~H^{+},~He,~He^{+},~He^{++},~e^{-},~H^{-},~H_{2},~H_{2}^{+}$ are self-consistently solved in the cosmological simulations. We make use of $\rm H_{2}$ photo-dissociating background UV flux implemented in the Enzo code. An external UV field of constant strength $\rm 10^{3}$ in units of $\rm J_{21}$ is used in the simulations. We presume that such flux is generated from a nearby star forming halo \citep{2008MNRAS.391.1961D} and is emitted by Pop III stars with a thermal spectrum of $\rm 10^{5}$ K. We include several cooling and heating mechanisms like collisional ionization cooling, radiative recombination cooling, collisional excitation cooling, H$_{2}$ cooling as well as $\rm H_{2}$ formation heating. The chemistry solver used in this work is a modified version of \cite{1997NewA....2..181A,1997NewA....2..209A}. We note that for columns above $\rm 10^{22}~ cm^{-2}$, the gas becomes optically thick to Lyman $\alpha$ photons, providing a potential complication for the further evolution. In fact, \citet{2006ApJ...652..902S} suggested that the latter may effectively stop the cooling and provide a transition to an approximately adiabatic regime. The latter was explored in detail by \citet{2010ApJ...712L..69S}, finding that at this point, additional processes become relevant, including the two-photon decay (2s-1s transition) and H$^-$ formation cooling. Effectively, the temperature evolution is then very close to the evolution obtained from optically thin Lyman $\alpha$ cooling. We also note that for a stellar spectrum of $\rm 10^{5}$ K $\rm H_{2}$ is mainly dissociated by the Solomon process while for the stellar spectrum of $\rm 10^{4}$ K the main dissociation route is $\rm H^{-}$ \citep{2010MNRAS.402.1249S,2011MNRAS.418..838W}. In principle, one would of course expect the presence of both contributions. The details of these processes however would not matter as long as H$_2$ is efficiently dissociated. \subsection{SGS turbulence Model} Due to the high Reynolds numbers relevant for astrophysical systems, it is not possible to resolve all scales down to the dissipative scale even with adaptive mesh refinement techniques. Turbulence cascades from coarser grids corresponding to large scales down to the center of the structure forming haloes without being properly accounted for. In engineering and many other disciplines of computational fluid dynamics subgrid scale turbulence models are used to represent the effect of unresolved turbulence on resolved scales. The unresolved turbulence has gained lot of interest in astrophysical simulations \citep{2008ApJ...686..927S,2009MNRAS.395.1875O,2009ApJ...707...40M}. To compute the unresolved turbulence on grid scales in our simulation, we use the subgrid scale (SGS) turbulence model by \citet{SchmNie06b}. This SGS model is based on a mathematically rigorous approach separating the resolved and unresolved scales, and connecting them via an eddy-viscosity closure for the non-linear energy transfer across the grid scale. The turbulent viscosity is given by the grid scale and the SGS turbulence energy, i.~e., the kinetic energy associated with numerically unresolved turbulent velocity fluctuations. The equations for compressible fluid dynamics are decomposed into resolved (large scales) and unresolved (small scales) parts using the filter formalism proposed by \citet{1992JFM...238..325G} in terms of density weighted quantities. Applying the filtering mechanism to the fluid equations and solving them in comoving coordinate system, we obtain \citep{SchmNie06b,2009ApJ...707...40M}: \begin{equation} {\partial \over \partial t} \langle \tilde{\rho} \rangle + {1 \over a } { \partial \over \partial x_{j}} \hat{u_{j}} \langle \tilde{\rho} \rangle = 0 \end{equation} \begin{dmath} { \partial \over \partial t} \langle \tilde{\rho} \rangle \hat{u_{j}} + {1 \over a }{\partial \over \partial x_{i}} \hat{u_{j}} \langle \tilde{\rho} \rangle \hat{u_{i}} = -{1 \over a } {\partial \over \partial x_{j}} \langle \tilde{p} \rangle + \langle \tilde{\rho} \rangle \hat{g_{i}^{*}} -{1 \over a }{\partial \over \partial x_{j}}\hat{\tau}(u_{i},u_{j}) \\ -{\dot{a} \over a } \langle \tilde{\rho} \rangle \hat{u_{j}} \end{dmath} \begin{dmath} {\partial \over \partial t} \langle \tilde{\rho} \rangle e_{res} + {1 \over a }{\partial \over \partial x_{j}} \hat{u_{j}} \langle \tilde{\rho} \rangle e_{res} = -{1 \over a } {\partial \over \partial x_{i}} \hat{u_{i}} \langle \tilde{p} \rangle - {\dot{a} \over a } (\langle \tilde{\rho} \rangle e_{res} + {1 \over 3 } \langle \tilde{\rho} \rangle \hat{u_{i}}\hat{u_{i}} \\ + \langle \tilde{p} \rangle ) + \langle \tilde{\rho} \rangle(\lambda + \epsilon)-{1 \over a } \hat{u_{i}} {\partial \over \partial x_{j}} \hat{\tau}(u_{i},u_{j}) \end{dmath} \begin{dmath} { \partial \over \partial t} \langle \tilde{\rho} \rangle e_{t} + {1 \over a } { \partial \over \partial x_{j}} \hat{u_{j}} \langle \tilde{\rho} \rangle e_{t} = \mathbb{D} + \Gamma - \langle \rho \rangle (\lambda + \epsilon) - {1 \over a } \hat{\tau}(u_{i},u_{j}) {\partial \over \partial x_{j}} \hat{u_{i}} \\ - 2{\dot{a} \over a } \langle \tilde{\rho} \rangle e_{t} \end{dmath} here a is the scale factor and the quantities with $\tilde{}$ are the comoving gas density, velocity and pressure etc. $\sigma_{ij}$ is the viscous stress tensor, q is turbulent velocity, $\lambda$ describes the effect of unresolved pressure fluctuations, $\epsilon$ accounts for the dissipation of kinetic energy and $\eta$ is the dynamic viscosity. Equations (1-3) are the mass, the momentum and the energy conservation equations. Equation (4) solves the evolution of SGS turbulent energy ($\rm e_{t}$). The quantities $\mathbb{D}$, $\lambda$, $\epsilon$, $\Gamma$ and $\hat{\tau}(\nu_{i}, \nu_{j})$ are unknowns and are computed in terms of closure relations, i.e., functions of the filtered flow quantities and the turbulent energy. The expressions for the closure terms are the following: \begin{equation} \mathbb{D} = { \partial \over \partial r_{i}} C_{\mathbb{D}} \langle \rho \rangle l_{\Delta}q^{2} { \partial \over \partial r_{i}} q \end{equation} \begin{equation} \hat{\tau}_{ij} = -2 \eta_{t} S^{*}_{ij} + {1 \over 3}\delta_{ij} \langle \rho \rangle q^{2} \end{equation} where \begin{equation} S^{*}_{ij} = {1 \over 2} \left({\partial \over \partial r_{j}} \hat{\nu_{i}} + {\partial \over \partial r_{i}} \hat{\nu_{j}} \right) + {1 \over 3} \delta_{ij}\langle \rho \rangle q^{2} \end{equation} \begin{equation} \lambda = C_{\lambda}q^{2}{\partial \over \partial r_{i}} \hat{\nu_{i}} \end{equation} \begin{equation} \epsilon = C_{\epsilon} {q^{3} \over l_{\Delta}} \left(1 +\alpha_{1} M^{2}_{t} \right) \end{equation} Further details and numerical validations of the applied closures are given in dedicated studies of \cite{SchmNie06b} and \cite{2009ApJ...707...40M} and are beyond the scope of this article. For our simulations, the coefficients of the model are calibrated against subsonic compressible turbulence simulations, comparable to the regime in our simulations \cite[see][]{SchmNie06b}. To apply the SGS model in cosmological AMR simulations, the method of adaptively refined large eddy simulations is employed \citep{2009ApJ...707...40M}. Since SGS turbulence energy depends on the grid scale, which varies on adaptive meshes, energy must be exchanged with the numerically resolved velocity field when the grid is refined or de-refined. This is achieved by assuming the Kolmogorov two-thirds law for the scaling of the turbulent velocity fluctuations. As long as the turbulence is subsonic and nearly isotropic locally, this is a reasonable assumption. In contrast, the method used by \citet{2011ApJ...733...88G} does not calculate the turbulence energy on the grid scale, but the energy associated with a characteristic length scale of buoyant bubbles. Our SGS model completely neglects gravity on unresolved length scales and assumes the turbulent cascade from larger resolved scales as the dominant source of unresolved turbulence \citep[for the effects of buoyancy on subgrid scales, see also][]{SchmNie06c}. Of course, this requires that turbulence is sufficiently resolved. The different resolutions considered in this study allow us to estimate biases from scale separation between the production of turbulence by gravity and the grid scale. \begin{figure*} \hspace{-4.0cm} \centering \begin{tabular}{c} \begin{minipage}{12cm} \includegraphics[scale=0.8]{Rescomp_profile1.ps} \end{minipage} \end{tabular} \caption{This figure shows the radially binned spherically averaged radial profiles for the halo C with and without SGS turbulence for different Jeans resolutions. The dashed lines show normal runs while the solid lines represent the cases with SGS turbulence as depicted in the legend. The upper left panel of the figure shows the density radial profiles. The temperature radial profiles are depicted the upper right panel. The bottom left panel of the figure shows the averaged total energy radial profiles. The averaged radial profiles of SGS energy are shown in the bottom right panel of the figure.} \label{fig} \end{figure*} \begin{figure*} \hspace{-4.0cm} \centering \begin{tabular}{c} \begin{minipage}{8cm} \includegraphics[scale=0.6]{Rescomp3.ps} \end{minipage} \end{tabular} \caption{The figure shows the density projections for the halo C with and without SGS turbulence. The state of simulations for the central 500 AU region of the halo is illustrated for various Jeans resolutions. The panels from top to bottom represent the Jeans resolution of 16, 32 and 64 cells respectively. The left panels present the normal runs while the cases with SGS turbulence are shown in the right panels.} \label{fig0} \end{figure*} \begin{figure*} \hspace{-4.0cm} \centering \begin{tabular}{c} \begin{minipage}{12cm} \includegraphics[scale=0.8]{ABHalos_profile.ps} \end{minipage} \end{tabular} \caption{The figure shows the radially binned spherically averaged radial profiles for halos A, B and C with and without SGS turbulence. The solid lines show the normal runs while the dashed lines represent the cases with SGS turbulence as shown in the legend. The upper left panel of the figure shows the density radial profiles. The temperature radial profiles are depicted in the upper right panel. The bottom left panel of the figure shows the total energy radial profiles. The averaged SGS energy radial profiles are depicted in the bottom right panel of the figure. } \label{fig1} \end{figure*} \begin{figure*} \hspace{-4.0cm} \centering \begin{tabular}{c c} \begin{minipage}{6cm} \includegraphics[scale=0.4]{H2I_profile.ps} \end{minipage} & \hspace{2cm} \begin{minipage}{6cm} \includegraphics[scale=0.4]{ELEC_profile.ps} \end{minipage} \end{tabular} \caption{The left panel of the figure shows the $\rm H_{2}$ abundance for three different haloes. The solid lines show the cases with no SGS turbulence and the dashed lines show the cases with SGS turbulence. The corresponding electron fraction is shown in the right panel.} \label{fig3} \end{figure*} \begin{figure*} \hspace{-4.0cm} \centering \begin{tabular}{c c} \begin{minipage}{6cm} \includegraphics[scale=0.4]{ABHJ640021_Density3.ps} \end{minipage} & \hspace{2cm} \begin{minipage}{6cm} \includegraphics[scale=0.4]{ABHJ640021_Vorticity3.ps} \end{minipage} \end{tabular} \caption{The left panel of this figure shows the density projections for the central 500 AU region of the halo. The images are shown for three different halos A, B and C from top to bottom respectively. The left side of left panel shows the normal runs while the right side depicts the cases with SGS turbulence. The right panel of the figure shows the density weighted vorticity projections corresponding to the density projections in the left panel. It can be noticed that halos with SGS turbulence model have more compact structures.} \label{fig5} \end{figure*} \begin{figure*} \centering \includegraphics[scale=0.8]{clumps-2.ps} \caption{The properties of the clumps are plotted against their masses in this figure. Diamonds show the data points for normal runs while data points for SGS turbulence cases are represented by triangles. Black, red and blue colors of the symbols represent the data for three different halos (i.e., A, B and C). The black solid lines are the fits to normal cases and the dot-dashed red lines are the fits to SGS cases. These clumps have sub-solar masses and are gravitationally unbound.} \label{fig6} \end{figure*} \section{Results} \subsection{Resolution comparison} In this study, we have performed simulations resolving the Jeans length by 16, 32 and 64 cells (hereafter called $\rm J_{16},J_{32},J_{64}$ respectively) and compared the results with SGS turbulence model. The properties of the halo for different Jeans resolutions with and without the SGS turbulence model are shown in figure \ref{fig}. The results are examined for the same peak density (i.e., $\rm 10^{-11}~gcm^{-3}$). The averaged density radial profiles for all the cases are depicted in the top left panel of figure \ref{fig}. The density profiles show almost $\rm R^{-2}$ behavior according to the expectation of a isothermal collapse. The bumps in the density profile for $\rm J_{16}$ and $\rm J_{32}$ cases indicate the presence of under-resolved density clumps in the vicinity of the central object. The overall density profiles agree well for different resolutions. The top right panel of figure \ref{fig} shows the averaged temperature radial profiles. It can be noticed that for all the cases the halo is heated up to its virial temperature and then cools by atomic line cooling. The temperature in the center of the halo remains about 7000 K. This shows that thermal properties of the halo are independent of resolution as well as SGS turbulence. The total energy for different Jeans resolutions and SGS turbulence cases is shown in the bottom left panel of figure \ref{fig}. The value of total energy increases at larger radii and becomes almost constant with a value of $\rm 3 \times 10^{12}~erg/g$ in the center of a halo. It is worth noting that these quantities are approximately converged, and do not depend strongly on the subgrid-scale model. Unlike in minihaloes, the results of such systems are thus likely more robust, as the thermal evolution is close to isothermal and less sensitive to minor changes in the dynamics. While the total energy is roughly the same for the three runs without SGS model, we see a non-monotonous behavior with resolution if the SGS model is applied. However, by considering the morphology of the halo in figure \ref{fig0}, it becomes clear that there are only little or no turbulent structure in the lower-resolution runs. For $\rm J_{16}$, turbulence is definitely under-resolved, while $\rm J_{32}$ could be an intermediate case. Consequently, the turbulent cascade cannot be sufficiently resolved in these runs. For the highest resolution case (i.e., $\rm J_{64}$), the structure inside the halo appears to be well resolved. The central region of a few hundred AU in size has become highly turbulent and clumpy. The turn-around in the trend for the total energy profiles indicates that Kolmogorov-like turbulence begins to be resolved for $\rm J_{64}$. This is plausible because the scale separation between the production of turbulence by gravity and grid scale is almost two decades, but the lower decade is strongly affected by numerical viscosity. Our results are also consistent with earlier studies \citep{2011ApJ...731...62F}. The profiles of the SGS turbulence energy for the different resolutions are shown in the bottom right panel of figure \ref{fig}. Since the grid scale decreases as the refinement levels become higher, the expectation based on the Kolmogorov scaling law would be that the SGS turbulence energy decreases toward the center and with the maximal resolution. However, this applies only to homogeneous turbulence. Since turbulence production tends to be stronger in the center of the halo, the resolution-dependence is compensated and yields a nearly constant SGS energy in the central region. The comparable plateaus of the SGS energy for $\rm J_{32}$ and $\rm J_{64}$ also indicate that the enhancement of SGS turbulence production in the $\rm J_{64}$ case compensates the smaller grid scale in comparison to $\rm J_{32}$. The SGS energy has a peak around radii of $\rm 5 \times 10^{21}$ cm ($\rm 2~kpc$), which coincides with the drop in the temperature profiles. The outward decrease of the SGS energy at larger radii shows that turbulence is mainly produced as a result of the collapse, but not by hydrodynamical instabilities outside of the halo. Furthermore, comparisons of the morphology of the halo for the three different Jeans resolutions reveal significant differences between the runs with and without SGS model (see figure \ref{fig0}). This is most obvious from the low-resolution runs that the SGS model is limited by the poorly resolved turbulence. However, we will concentrate on the highest-resolution case for quantitative comparisons in the next section. \subsection{Study of different Haloes} \subsubsection{Global Dynamics of collapse} We have performed six cosmological simulations for three different haloes (named A, B and C) with a constant strength $\rm J_{21}=10^{3}$ of the $\rm H_{2}$ photo-dissociating radiation field. The masses of the haloes and their collapse redshifts are listed in table \ref{table1}. The results obtained from cosmological simulations conducted in this work are presented in the following subsections. The density fluctuations collapse under the gravitational instability as they decouple from the Hubble flow. These small fluctuations merge with each other to form larger haloes in accordance with the standard paradigm of structure formation. In the early phases of the collapse gas falls in the dark matter potentials and gets shock-heated during the nonlinear evolution phase. Gravitational energy of the halo is continuously transferred to kinetic energy of the gas and dark matter during the course of virialization. Here we report the study of three haloes of different masses but with same Jeans resolution to compute the variation from halo to halo. Again, we compare our results with SGS turbulence model. The averaged density radial profiles for the haloes A, B and C with and without SGS model, centered at the densest cell are shown in the upper left panel of figure \ref{fig1}. The maximum density in our simulations is a few $\rm 10^{-11}~gcm^{-3}$. It can be seen from the figure that all cases follow almost $\rm R ^{-2}$ behavior which would be expected from a isothermal collapse. There is a bump in the density profile of halo A with SGS turbulence model which is an indication of fragmentation. In the very central region the density profile is flat which corresponds to the local Jeans length in all cases. These density profiles are comparable to the previous studies \citep{2002Sci...295...93A,2008ApJ...682..745W,2009Sci...325..601T,2012ApJ...745..154T}. The average radial profile of total energy is shown in the bottom left panel of figure \ref{fig1}. The total energy of the system is a few times $\rm 10^{12}~erg~g^{-1}$ for all the cases. The figure shows the total energy for three different halos with and without SGS turbulence is converged. The increase in the total energy radial profile towards the smaller radii is due to gas infall in the center of halo. The specific subgrid scale energy for the three different halos is shown in the bottom left panel figure \ref{fig1}. At larger radii, the build-up of turbulent energy increases sharply because of the turbulence cascade and then gets saturated at smaller radii due to the enhanced turbulence dissipation rate. This evolution of SGS energy is according to the expectations of large eddy simulations \citep{2009ApJ...707...40M}. \begin{table} \begin{center} \caption{The halo masses and their collapse redshifts are listed in this table.} \begin{tabular}{cccccc} \hline \hline Model & Mass & Collapse redshift \\ & $\rm M_{{\odot}} $ & z \\ \hline \\ A & $\rm 8.06 \times 10^{6}$ & 11.9 \\ B & $\rm 4.3 \times 10^{6}$ & 11.3 \\ C & $\rm 3.2 \times 10^{7}$ & 14.1 \\ \hline \end{tabular} \label{table1} \end{center} \end{table} \subsubsection {Thermodynamics} The thermal evolution of the gas for different halos with same Jeans resolution is shown in the right panel of figure \ref{fig1}. During the process of virialization, the gas is heated up to its virial temperature (i.e., $\rm \ge 10^{4}$K) and subsequently cools by Lyman alpha radiation. Consequently, gas collapses almost isothermally with temperatures around 8000 K. The temperature profiles for all the cases with and without SGS turbulence are similar. There is no significant turbulent heating for SGS turbulence cases as seen in \cite{2009ApJ...707...40M} due to highly efficient atomic line cooling at these temperatures. The dissimilarities in the temperature profiles at larger radii appear due to the difference in the halo masses. The ubiquity of intense Lyman Werner radiation photo-dissociates the molecular hydrogen and $\rm H_{2}$ cooling remains suppressed. Our results are in agreement with previous studies \citep{2003ApJ...596...34B,2008ApJ...682..745W,2011A&A...532A..66L,2012A&A...540A.101L} and according to the expectation of theoretical models. The H$_{2}$ abundance is shown in the left panel of figure \ref{fig3}. It can be noticed that the H$_{2}$ fraction increases at lower densities due to the rise in electron abundance during the non-linear phase of the collapse. At intermediate densities, the H$_{2}$ abundance becomes constant as gas cools, recombines and remains neutral with a constant temperature around 8000 K. The presence of sharp spikes in the H$_{2}$ fraction is due to the shocks occurring at the central densities due to collisional dissociation. In general, the H$_{2}$ fraction is lower than the universal value (i.e., $\rm 10^{-3}$). Therefore, the contribution of H$_{2}$ cooling in the thermal evolution of the haloes studied here is negligible. The electron abundance corresponding to the H$_{2}$ fraction is depicted in the right panel of fig \ref{fig3}. It can be seen that the electron abundance is correlated with the H$_{2}$ fraction. At densities above 1 $\rm cm^{-3}$, the electron fraction increases because of virialization shocks and then continues to decline as gas becomes neutral. Small wiggles in the central electron fraction are triggered by shocks. \subsubsection {Halo structure} The state of the simulations with and without the SGS turbulence at the collapse redshift is illustrated by the density projections in the left panel of figure \ref{fig5}. Significant changes in the morphology of haloes are found in the presence of SGS turbulence in all haloes (i.e., A, B and C). It can be noted that haloes are highly turbulent and clumpy in both cases. These effects were not seen in the earlier studies due to the poor Jeans resolution. Our results confirm that one needs to resolve the Jeans length with at least 32 cells or higher to capture turbulent velocity fluctuations \citep{2011ApJ...731...62F,2012ApJ...745..154T}. Overall, halos in simulations with SGS turbulence are more compact and denser than their counterparts. The latter is expected due to the presence of an additional viscosity term. As demonstrated above, these haloes have a similar thermal evolution but the changes in the morphology arise as a consequence of unresolved subgrid scale energy computed via the SGS turbulence model and show substantial variation from halo to halo. The SGS energy is about 10 \% of total energy budget. Its effect is particularly enhanced on small scales, yielding rather different morphologies in the presence of the subgrid-scale model. The structure of the halo ''A'' without SGS turbulence clearly shows that dense clumps are very well separated from each other and may lead to the formation of a binary in this case. The further evolution of the simulations becomes computationally very demanding as the Jeans mass keeps decreasing. Here, we stopped our simulations after reaching the maximum refinement level. We plan to explore the further evolution of at least one halo in a companion paper. To determine the presence of turbulent velocity fluctuations, we have computed the fluid vorticity (i.e., $\nabla \times v$). Density weighted projections of the vorticity squared centered at the densest point are depicted in the right panel of figure \ref{fig5}. In the center of the haloes, large regions with high values of vorticity indicate the ubiquity of high turbulent energy. It is also noted that high values of the vorticity are correlated with the dense regions of the haloes. These vorticity plots further suggest the absence of coherent structures. The amount of vorticity is higher compared to the SGS turbulence cases because of higher turbulent dissipation rates in SGS turbulence cases. \subsubsection{ Properties of clumps} In order to quantify the properties of the clumps found during visual inspection, we have employed the clump finder of \citet{1994ApJ...428..693W}. The properties of the clumps with and without SGS turbulence model for three different haloes (A, B and C) with power law fits are shown in figure \ref{fig6}. The top panel of figure \ref{fig6} depicts that clumps are not gravitationally bound as their masses are smaller than the Jeans mass and the number of clumps is generally higher in no SGS cases. The ratio of clump mass to Jeans mass shows a power law behavior (i.e., $\rm M/M_{J} \propto M^{1.3}$). It is interesting that similar trends have been found in different studies exploring clumps in molecular clouds \citep{2009MNRAS.398.1082B}. We find that in simulations with the SGS turbulence model, the clumps have slightly higher masses as shown by the fit (red line in figure \ref{fig6}) and the number of low mass clumps is reduced compared to the no SGS cases. This is likely an effect of the turbulent viscosity, which provides an additional diffusion mechanism that counteracts the formation of low-mass clumps. The thermal properties of the clumps are depicted in figure \ref{fig6}, showing that the clumps in simulations with SGS turbulence have almost the same temperature as those in the standard setup. The density of the clumps plotted against mass shows a bimodal distribution with clumps sitting at higher densities following a linear relation with $\rm M$. The notable difference is that clumps in SGS turbulence case have higher densities and the power law sharply drops for lower densities. The latter suggests that only the high-mass clumps manage to form in the presence of an additional turbulent viscosity, providing a distribution with more massive clumps on average. The velocity dispersion in the clumps increases with the mass and follows a $\rm M^{0.38}$ power law. Again, it is seen that clumps with SGS turbulence have lower values of dispersion velocity but roughly follow the same trend. The radii of the clumps comply a $\rm M^{0.8}$ growth and clumps with SGS turbulence model have larger radii in comparison with no SGS cases. In the last panel of figure \ref{fig6}, we show that these clumps are well resolved at least by $\rm 10^{5}$ cells. \section{Discussion and Conclusions} We have conducted high resolution cosmological simulations using the AMR code Enzo for three different halos with $\rm T_{vir} \geq 10^{4}$K irradiated by a constant strength of photo-dissociating background UV flux. In one set of simulations, we used the subgrid scale (SGS) turbulence model proposed by \citet{SchmNie06b} and \citet{2011A&A...528A.106S} to compute the kinetic energy of numerically unresolved turbulence and the associated stresses on resolved length scales. For comparison, we run these simulations also without the SGS model. Since a high dynamical range is crucial, we use two initial nested grids and insert up to 27 additional refinement levels during the course of simulations, corresponding to an effective resolution down to sub AU scales. To investigate resolution effects, we applied refinement at 16, 32 and 64 cells per Jeans length. The results from three haloes with different masses were examined to study the variation from halo to halo for a fixed resolution. The main conclusions from this study are the following: \begin{itemize} \item The global properties of the halo, in particular the radial profiles, are converged and can be used as a robust input for direct collapse models. \item Turbulent structures are observed for a Jeans resolution of at least $\rm \geq 32$ cells. \item The morphology of the halo and its clump properties are strongly influenced by taking into account SGS turbulence and typically more compact. \item The clump properties (i.e., $\rm M/M_{J}$, velocity dispersion) show a power law behavior against clump masses. \end{itemize} The gas in the atomic cooling halos is heated up to its virial temperature where Lyman alpha cooling comes into play and cools the gas down to 8000 K. The presence of an intense Lyman Werner UV radiation field of $\rm 10^{3}$ in units of $\rm J_{21}$ photo-dissociates the $\rm H_{2}$ molecules via the Solomon process. We employed grid resolutions of $16$, $32$ and $64$ cells per Jeans length to explore the convergence of global properties as well as the local morphology. It is important to note that the radial profiles of density, temperature and total energy are approximately converged, implying more robust results than previously reported for minihalos \citep{2012ApJ...745..154T}. We attribute the latter to the thermodynamics of these halos, which are considerably more robust in the presence of strong H$_2$ photodissociation \citep{2010ApJ...712L..69S}. In this case, the temperature evolution remains very close to isothermal, making it rather insensitive to local changes in the dynamics. With such a fixed thermal pressure, the resulting evolution during the collapse is therefore considerably more robust. We note that SGS turbulence does not have a strong impact on thermodynamical properties, again as a result of efficient Lyman $\alpha$ cooling. The typical unresolved fraction of the turbulence energy is $\rm \simeq 10 \%$ of the total energy. Our results demonstrate with the highest resolution simulations that atomic cooling halos become highly turbulent. We computed the evolution of three different halos and found that the radially averaged properties are very similar for three different halos, but the morphology of the haloes varies considerably. The intense vorticity inside the haloes demonstrates the absence of coherent structures in fully turbulent regions. However, since the non-linear coupling between gravity and turbulence converts gravitational potential energy into kinetic energy, an inertial range of hydrodynamical turbulence can only exist on length scales smaller than the Jeans length. For this reason, a sufficient range of length scales smaller than the Jeans length must be resolved to observe a turbulent cascade with an inertial sub-range of the Kolmogorov type (for low compressibility) in simulations. This pushes numerical simulations to their limits because an extremely high dynamical range is required. In our simulations with 16, 32 and 64 cells per Jeans length, we find significant differences concerning the central morphologies and the amount of turbulent structures. The latter confirms that turbulence is only marginally resolved even at the highest resolution \citep[see also][]{2011ApJ...731...62F}. This is also reflected by the radial profiles of the turbulent energy in the simulations with SGS model, for which no clear convergence trend is found. To verify convergence, higher resolution $\rm \geq$ 64 cells per Jeans length should be performed in the future. Based on the notion of large eddy simulations, one would naively expect that the application of an SGS model reduces the range of length scales that has to be resolved. However, large eddy simulations are applicable only if the turbulent cascade is partially resolved, i.~e., at least by a decade in scale space. On top of that, numerical dissipation typically necessitates an additional decade. In simulations of turbulent collapsing halos, this corresponds to scale separation between the the Jeans length, at which energy is injected by gravity, and the grid scale. Apart from that, lack of local isotropy at low resolutions poses a problem because we use a constant coefficient for the production of SGS turbulence energy by shear. This problem could be addressed, for example, by a localized SGS model with varying coefficients \cite[see][]{SchmNie06c}. However, this method has not been applied yet in AMR simulation. Overall, it is important to note that, while a subgrid-scale model does not yield convergence at low resolution, it certainly does improve the solution once a sufficiently high resolution is reached, such that its central assumptions are fulfilled. In fact, it is clear that direct numerical simulations resolving the turbulence over a sufficient range of scales cannot be pursued in the near future. Obtaining robust astrophysical results therefore requires both high numerical resolution, as well as subgrid models to account for effects still below the grid scale. On the basis of the current simulations, we can already draw relevant conclusions about the effects of numerically unresolved turbulence for the Jeans resolution of 64 cells. By comparing the results with and without SGS cases, it was found that the morphology of the halos obtained in the simulations with SGS model is significantly different from their counterparts. In general, their central gas distributions are denser and more compact compared to the none-SGS cases. The latter provides an indication that larger accretion rates onto the central object can be expected due to the additional turbulent viscosity. In one case, the density structure of the halo shows a bimodal distribution if no SGS model is applied, which might result in the formation of a binary. Such structural implications need to be addressed employing numerical methods like sink particles or a pressure floor. It is thus worth exploring whether statistical differences can also be obtained for gravitationally bound clumps, and how much their accretion is enhanced via turbulent viscosity. We noticed that structures become compact in the presence of SGS turbulence. This may have important implications for the accretion of mass to the central object, potentially favoring the higher accretion rates. In our simulations, we observed the formation of turbulent structures for resolutions of $\rm \geq 32$ cells per Jeans length and no accretion disk at this stage of the collapse. We do not follow the evolution of these haloes for longer dynamical times and therefore cannot make statements about the final fate of these haloes. According to theoretical predictions, it is likely that formation of disk may take place. Numerical investigations of the direct collapse scenario thus need to employ a sufficiently high numerical resolution as well as a turbulence subgrid model to determine realistic accretion rates. \section*{Acknowledgments} The simulations described in this work were performed using the Enzo code, developed by the Laboratory for Computational Astrophysics at the University of California in San Diego (http://lca.ucsd.edu). We thank Matt Turk, Robi Banerjee, John Wise and Enzo developers for helpful discussions. This work was supported from the SFB~963 (project A12) {\em Astrophysical Flow Instabilities and Turbulence}. We also acknowledge the funding support from German Science Foundation. DRGS thanks for funding from the Deutsche Forschungsgemeinschaft (DFG) in the Schwerpunktprogramm SPP 1573 “Physics of the Interstellar Medium” under grant SCHL 1964/1-1. The simulation results are analyzed using the visualization toolkit for astrophysical data YT \citep{2011ApJS..192....9T}.
\section{Introduction} Spin ordering, and often orbital ordering, is normally unambiguous, as these properties are subject to direct observation by magnetic and spectroscopic measurements, respectively. Charge ordering (CO) and the actual charge of an ion is rarely measured directly, and the formal charge of an ion in the solid state can be a point of confusion and contention. Valence, oxidation number, and formal charge are concepts borrowed from chemistry, where it is emphasized they do not represent actual charge\cite{chem1,chem2} and have even been labeled hypothetical.\cite{chem1} As the interplay between spin, charge, orbital, and lattice degrees of freedom become more closely watched\cite{DIKGAS} and acknowledged to be a complex phenomenon, disproportionation and CO have become entrenched as the explanation of several high profile metal-insulator transitions (MIT). The possibility that CO in the charge transfer regime is associated with the oxygen sublattice, with negligible participation of the metal, has been raised\cite{mizokawa} and considered as an alternative.\cite{Mazin} Charge density is a physical observable of condensed matter, and the desire to assign charge to atoms has evident pedagogical value, so theoretical approaches have been devised to share it amongst constituent nuclei. Mulliken charge population, which socializes shared charge (divides it evenly between overlapping orbitals) is notoriously sensitive to the local orbital basis set that is required to specify it. Born effective charges are dynamical properties and are often quite different from any conceivable formal charge or actual charge. Integrations over various volumes have been used a great deal, but dividing the static crystal charge density into atomic contributions is, undeniably, an ill-defined activity. A possibility that has not been utilized is that, taking $3d$ oxides as an example, there is a directly relevant metric that is well defined: the $d$ occupation $n_d$. This quantity is in fact what the physical picture of formal charge or oxidation state brings to mind. $3d$ cations, in their various environments and charge states, have maxima in their spherically averaged radial density $\bar{\rho}(r)$ in the range 0.6-0.9 $a_o$. At this short distance from the nucleus, the only other contribution to the density is the core contribution, which can be subtracted out and is unchanged during chemical processes or CO. Most relevant to the understanding of CO-driven transitions and disproportionation is the (actual or relative) {\it difference} in $3d$ occupations $\Delta n_d$, which is given directly, {\it without any integration}, by the difference in the radial $3d$ densities at their peaks, where there are no competing orbital occupations to confuse charge counting. This specifically defined $3d$ occupation differences provides a basis for building a faithful picture of CO and of characterizing formal valence differences more realistically. We consider our computational results\cite{LAPW,wien} for a selection of systems, then discuss some of the implications. \underline{La$_2$VCuO$_6$} (LVCO) is a double perovskite compound providing a vivid and illustrative example. Our earlier study\cite{vcu} revealed two competing configurations for the ground state. Using conventional identifications, one is the V$^{4+}$ $d^1$, Cu$^{2+}$ $d^9$ magnetic configuration (with bands shown in Fig. \ref{VCuBands}) identified as such because (1) there is one band of strong V $d$ character occupied and one band of strong Cu $d$ character unoccupied, and (2) the moments on both V and Cu, 0.7 $\mu_B$, are representative of many cases of spin-half moments reduced by hybridization with O $2p$ orbitals. The other configuration is the nonmagnetic $d^0-d^{10}$ band insulator: all Cu $d$ bands are occupied, all V $d$ bands are unoccupied -- a conventional ionic band insulator in all respects. The identification of formal valence (or oxidation state) is crystal clear. \noindent \begin{figure}[!htb] \includegraphics[width=\columnwidth,angle=0]{la2vcuo6_d1d9_Vmajority.eps} \includegraphics[width=\columnwidth,angle=0]{la2vcuo6_d1d9_Cuminority.eps} \caption{(Color online) Top: bands near the Fermi energy/bandgap in the $d^1 - d^9$ magnetic, nearly Mott insulating, configuration of La$_2$VCuO$_6$. The $d_{xy}$-up band is correlation-split off from the other two $t_{2g}$ bands and fully occupied. Bottom: the Cu fatbands for the same system, showing one unoccupied Cu minority $d_{x^2-y^2}$ band correlation-split from the $d_{z^2}$ band. The other $d$ bands fall outside this energy range. \label{VCuBands} } \end{figure} The radial charge densities of V and of Cu for both configurations reveal an unsettling feature: the actual $3d$ occupations $n_d$ of each of these V and Cu ions are {\it identical for both configurations}, in spite of the unit difference in their formal charges. (Identical in this paper means to better than 0.5\% $\sim 0.01 e^-$, in terms of the differences of charge density at their peaks.) Thus ions with {\it no real difference} in $3d$ occupation can behave as if they comprise charge states differing by unity. Changes in spin-orbital occupations, which quantify spin, orbital, and charge differences between the two states, can be quantified by the LDA+U spin-orbital occupations. For the V $d^1$ $d_{xy}$ (Jahn-Teller split) orbital, the majority-minority difference is 0.70, which accounts for all of the moment. The difference of 0.65 between $d_{xy}$ and each of the other $t_{2g}$ characterizes the Jahn-Teller distortion. The increase in charge of the $d_{xy}$ orbitals (both spins), 0.55, compared to the $d^0$ state, is absorbed more or less uniformly from all other (nominally unoccupied) spin orbitals. Similarly for Cu, the $d^9$ hole results from a difference of charge in the minority $d_{x^2-y^2}$ orbital of 0.6, with the other hole charge being distributed nearly uniformly over the other nine (nominally but not actually fully occupied) spin-orbitals. In both cases the moment arises entirely from the single magnetic orbital as the simple picture would suggest, while all other orbitals are unpolarized. This happens, conspicuously, with {\it no change} in $n_d$ for either V or Cu. Charge is redistributed to one orbital from the others, and strongly spin-imbalanced within that orbital. Even with insulators with ``obvious'' charge states, $3d$ orbital occupations can range over the values [0,1]. We look at additional cases before addressing some of the implications. \underline{Rare earth (${\cal R}$) nickelates ${\cal R}$NiO$_3$} display a first order structural and MIT of great current interest. The $Pbnm$ (GdFeO$_3$ structure) $\rightarrow$ $P2_1/n$ transformation results in a large Ni1O$_6$ and a small Ni2O$_6$ octahedron, with Ni-O distances of 2.015$\pm$0.015 \AA~ and 1.915$\pm$0.025 \AA, respectively, that are not otherwise strongly distorted; see the inset of Fig.~\ref{Radial}. At a temperature that varies smoothly from 600K to 300K with increasing ${\cal R}$ ionic radius, the resistivity of these nickelates drops sharply.\cite{Garcia1992,Torrance1992} We focus on YNiO$_3$; with its small ionic radius, it is one of the more strongly distorted members, and the resulting narrowed bandwidths make it more prone to strong correlation and CO tendencies.\cite{Mazin} Structural changes at the MIT have been studied extensively,\cite{Garcia1992,ynio3,Alonso,Alonso1999,I.Vobornik1999} which together with x-ray absorption spectral splittings\cite{Staub,Piamonteze,Medarde2009} have been interpreted in terms of charge disproportionation (or CO) 2Ni$^{3+} \rightarrow$ Ni$^{3+\delta}$ + Ni$^{3-\delta}$, with $\delta \approx$ 0.3 for YNiO$_3$.\cite{Staub} This MIT in the nickelates has been recognized as paradigmatic by theorists. Mizokawa et al. modeled this system\cite{Miza} with a multiband Hartree-Fock model in the charge-transfer regime and found evidence for CO on the {\it oxygen} sublattice for larger ${\cal R}$ cations, but concluded that YNiO$_3$ was representative of a CO transition on the Ni sites. Mazin {\it et al.}\cite{Mazin} surveyed the competition between Jahn-Teller distortion of the $d^7$ ion and CO and also concluded that YNiO$_3$ is a prime example of a CO $d^6 + d^8$ system. Lee {\it et al.} have investigated\cite{Balents} a two band model for this system with a CO interaction in mean field, emphasizing CO effects. On the other hand, Yamamoto and Fujiwara\cite{Fujiwara2002} reported a very small ($\sim$0.03 $e^-$) density functional based charge difference. \noindent \begin{figure}[!htb] \begin{center} \includegraphics[width=\columnwidth,angle=0]{YNiO3_radial.structure.eps} \end{center} \caption{(Color online) Radial charge density (upper curve) of YNiO$_3$ for $Pbnm$ Ni and $P2_1/n$ Ni1 and Ni2, showing there is no difference at the peak, which reflects the $3d$ occupation of the ion; a small difference shows up near the sphere boundary. The spin decompositions give easily visible differences. The vertical lines at the bottom right indicate conventional Ni$^{4+}$, Ni$^{3+}$, and Ni$^{2+}$ ionic radii, which have no relation to the (unvarying) $3d$ occupation. Inset: Structure of the broken symmetry $P2_1/n$ phase, showing the rotation in the $a-b$ plane and tilting along the $c$ axis of the NiO$_6$ octahedra (Ni is inside) and the ($\pi,\pi,\pi$) ordering of the Ni1 and Ni2 octahedra. \label{Radial} } \end{figure} For the assumed (for simplicity) ferromagnetic order the calculated Ni1 and Ni2 moments are 1.4 and 0.65 $\mu_B$ respectively for YNiO$_3$ and several other members of this class, so these values are not sensitive to the magnitude of the distortion. They coincide with the values obtained from neutron diffraction,\cite{ynio3} 1.4(1) and 0.7(1) $\mu_B$ respectively, in the magnetically ordered phase. It is intriguing that the same moments were obtained in fully relaxed LaNiO$_3$/LaAlO$_3$ monolayer superlattices.\cite{Blanca} The $3d$ occupations, obtained as above directly from the maximum in the radial charge density plots in Fig. \ref{Radial}, are identical for Ni1, Ni2, and the single Ni site in the high temperature phase: there is no $3d$ charge transfer, or disproportionation, across the transition. The majority and minority radial densities and integrated charges of course differ (see Fig. \ref{Radial}) as they must to give the moment, but the total $3d$ occupation is inflexible. This constancy of the $3d$ occupation across the transition, and equality for Ni1 and Ni2, is inconsistent with microscopic disproportionation. To illustrate the spin-orbital spectral density redistribution, the projected densities of states are shown in Fig. \ref{DOS}. All $t_{2g}$ states are filled and irrelevant. The $e_g$ spectral distribution is non-intuitive: weight from -5 eV spin-down is transferred to -1 eV spin-up. The majority $e_g$ states just below the gap are strongly Ni1 in character, while the unoccupied bands just above the gap are primarily Ni2. Such behavior is expected for different charge states, similarly to the behavior in LVCO above; however, the total $3d$ occupation is identical. The main differences between Ni1 and Ni2 show up in the {\it unoccupied} $e_g$ states: the Ni1 spin splitting is 3.5 eV, a reflection of the on-site repulsion that opens the Mott gap in the majority $e_g$ states, rather than Hund's exchange splitting. The origin of the Ni2 moment is murky, not identifiable with any occupied spectral density peak. Note that in a Ni$^{2+}$ $+$ Ni$^{4+}$ CO picture, Ni2 would be nonmagnetic. Not only is this calculated behavior not consistent with a CO picture, it involves redistribution not accounted for in any simple model. In spite of identical $3d$ charges, the Ni1 and Ni2 core energies differ by up to 1.5 eV. \noindent \begin{figure}[!htb] \begin{center} \includegraphics[width=\columnwidth,angle=0]{CO_dos.eps} \end{center} \caption{(Color online) Spin-decomposed Ni $t_{2g}$ and $e_g$ density of states for the Ni1 and Ni2 ions in the insulating $P2_1/n$ broken symmetry phase. The hashed regions illustrate the spectral origin of the enhanced moment of Ni1 relative to Ni2. The horizontal arrows illustrate the large difference in spin splittings, the result of the combination of Hund's coupling and Coulomb $U$= 5.7 eV. \label{DOS} } \end{figure} \underline{CaFeO$_3$}, another perovskite that displays the same $Pbnm \rightarrow P2_1/n$ structural change at T$_{MI}$ as the nickelates, is also explained\cite{Takano} in CO language that invokes the unusually high (penta)valent state Fe$^{5+}$. Analogously to YNiO$_3$, we obtain identical $3d$ occupations for Fe1 and Fe2 ions. Quantum chemical embedded cluster calculations\cite{QuantumChemical} and LDA+U studies\cite{LDA+U,saha,matsuno} had noted that the Fe charge in both ``disproportionated'' sites differed little, but neither quantified the occupation as we have for YNiO$_3$ and CaFeO$_3$. The pentavalent state of Fe has most often been identified from M\"ossbauer isomer shift data, but Sadoc {\it et al.}\cite{QuantumChemical} concluded the difference in isomer shift is primarily a measure of the covalency (Fe-O distance) rather than any real charge on Fe. \underline{AgNiO$_2$}, a triangular, magnetically frustrated lattice compound with nominal Ni$^{3+}$ ions, undergoes a structural transition at 365 K although remaining metallic.\cite{wawrzy1,wawrzy2,chung,pascut} Three inequivalent Ni sites arise, with a high spin Ni1 ion in an enlarged octahedron and two low spin Ni2, Ni3 = Ni2,3 ions in small octahedra. Based on the structural changes (which were quantified in terms of bond valence sums), the magnetic moments, and resonant x-ray scattering that confirms a calculated $\sim$1 eV difference in core level energies between Ni1 and Ni2,3, this transition has been welcomed as the first realization of such a highly unusual 3$e_g^1 \rightarrow e_g^2 +2 e_g^{0.5}$ type of CO. Furthermore, using the charge difference per unit core level splitting of 0.66 $e$/eV led to an inferred charge disproportionation of $\sim$1.65$e$, {\it i.e.} Ni1$^{2+}$ + 2 Ni2,3$^{3.5+}$. We have reproduced several of the first principles results\cite{wawrzy1,pascut} that were used to support CO. The calculations give a large moment ($> 1 \mu_B$) on high-spin Ni1 and very weak moments ($\sim$0.1 $\mu_B$) on low-spin Ni2,3 ions. We find, as in the cases above, that $n_d$ for the three sites are {\it identical}. Moreover, our calculated core level differences, 0.6-0.8 eV, are roughly consistent with reported values\cite{pascut} ($\sim$1 eV). \underline{V$_4$O$_7$} represents another oxide currently explained by a CO-driven MIT. It is structurally more involved, but first principles calculations of moments and geometries again have produced several results corroborating the experimental data\cite{v4o7,Hodeau} and were used to support CO into V$^{3+}$ and V$^{4+}$ charge states on specific sites. As in the instances above, we find no differences in $n_d$: the occupations are indistinguishable. The site energy differences, measured by differences in $1s, 2s, 2p$ core levels, differ by 0.9-1.2 eV for two sites, similar to the nickelates. The interplay of orbital order, structural distortions, and possible spin-singlet formation of half of the V ions provide a rich array of degrees of freedom, which can operate without need for disproportionation. \underline{Implications.} We have established that, for several instances of CO transition systems as well as for the two self-evident charge states of LVCO, there is no difference in the $3d$ occupations for the different ``charge states'' that have been used to categorize their behavior. Such identification is possible because a choice of a region for integration is avoided; the peak charge region rather than tails of orbitals are used in the identification. This finding of constancy sharpens several reports of ``small charge differences'' between differing charge states ({\it viz.} Luo {\it et al.}\cite{Luo} for doped manganites; Haldane and Anderson\cite{haldane} in a multi-orbital Anderson model, and Raebiger {\it et al.}\cite{lany} from DFT calculations for TM impurities in semiconductors; Yamamoto and Fujiwara\cite{Fujiwara2002} and also Park {\it et al.}\cite{rnio_millis} for nickelates). We see two primary implications: (1) the conceptual basis underlying a substantial aspect of transition metal physics is misleading, and (2) modeling of structural and electronic transitions has, at least in several conspicuous cases, incorporated the wrong mechanisms by invoking inactive degrees of freedom. Actual cases of CO very likely do exist, but the burden of proof has shifted. For these CO systems, the constancy of $n_d$ suggests that $U_d$ is too large to allow change in occupation $n_d$ in or near the ground state (in the cases we discuss, and similar ones). In insulators the charge is more physically pictured in terms of (fully occupied) Wannier functions (WFs) than in terms of ambiguous populations of atomic orbitals, making them appear to be inviting. However, WFs are far from unique and, like molecular orbitals, WFs contain charge that cannot objectively be assigned to one atom or another, so a WF viewpoint is not promising. A broader implication is that modeling of coupled structural and electronic transitions in terms of charges\cite{Mazin,Balents} from atomic-like orbitals must be treated with caution: charge fluctuations in these systems are too high in energy to comprise a relevant degree of freedom. The important energy differences are characterized in terms of differences in hopping amplitudes, anion-cation distances, and (not recognized in most models) resulting changes in site energies, as well as very important Hund's rule energies. Models that try to parametrize (for example) Ni1-Ni2 differences by on-site charge will not be treating the relevant microscopic degrees of freedom. CO on the oxygen sublattice\cite{mizokawa,Mazin} may also be problemmatic. Charge states of ions serve to specify the occupations of spin-orbitals. The essential degrees of freedom in determining this popular characterization, which professes to be quantitative, are the spin-orbital occupations, not as determined from the (real) density matrix but rather from the site symmetry, crystal symmetry, and the local moment. The LVCO example illustrates vividly how two different charge states, for both {\it highly charged} V and {\it moderately charged} Cu, can be represented by integer occupation of different numbers of orbitals while there is no change in $n_d$. ``Charge state'' projects onto integrally occupied orbitals, while the distribution of real charge is strongly non-integral and often non-intuitive. These projections are backed up by the number of occupied spin-polarized bands (an integer), by the (discrete) local symmetry (JT distortion), by the local moment (with its quantization smeared by hybridization), and by the atomic radii, but each one of these characterizations is extremely flexible with a given amount of $3d$ charge. More specifically to CO systems, the ionic environment in the high symmetry phase requires closer scrutiny. In both the nickelates and in V$_4$O$_7$ there is evidence of distinct metal sites {\it above} the transition, in the (on average) symmetric phase, and the structural similarities of CaFeO$_3$ to $R$NiO$_3$ suggest similar behavior there. For nickelates, x-ray absorption spectra\cite{Medarde2009,Piamonteze} reveal that local signatures of Ni1 and Ni2 sites persist continuously across the MIT, and both sites also remain when driven across the phase boundary by pressure.\cite{Ramos} As we have shown, the coordination alone ({\it i.e.} with identical $n_d$) accounts for on-site energy differences of $\sim$1 eV in spectra that have often been used to support disproportionation. The MITs in some of these materials may be primarily order-disorder type; the onset of long-range order in nickelates results in carrier localization and gap formation, ergo a MIT but one unrelated to CO. We propose therefore that ``charge order'' should be used as the name, hence the interpretation, of a phase transition only if an objective, relevant charge difference is the likely mechanism; otherwise, the underlying mechanisms should be identified. Formal developments may be useful; for example, Jiang {\it et al.} have provided a specification\cite{jiang} of integer charges in an insulator that they propose as oxidation states (which are identical to charge states in metal oxides.) Based on integration over a configuration space path of the dynamic Born effective charge, their expression assigns (in principle) an integer charge to each atom in any insulator. Notably, their specification does not refer to $3d$ charge explicitly and furthermore depends explicitly on dynamical effects (electron response to ion motion). Also, many CO interpretations only hold water if the supposed charge difference $\pm \delta$ is much smaller than unity ($\delta \sim$ 0.3 for the nickelates). More experience will be needed to learn how best to interpret their definition. Work at UC Davis was supported by DOE grant DE-FG02-04ER46111. V.P. acknowledges support from the Spanish Government through the Ram\'{o}n y Cajal Program.
\section{Introduction} More than a hundred young massive stars, mostly Wolf Rayet/O and B types, have been identified within a distance of $\sim 0.5{\rm ~pc}$ from the massive black hole (MBH) in the Galactic center \citep[GC;][]{Gillessenetal09,LuJ09,Bartko10}. These young stars are empirically divided into two groups: (1) the majority of the young stars at a distance $\sim 0.04$--$0.5{\rm ~pc}$ from the MBH are located on coherent disk-like structures, i.e., the clockwise rotating stellar (CWS) disk and the possible counterclockwise rotating stellar (CCWS) disk \citep[e.g.,][]{LB03,Paumard06,LuJ09}; and (2) the young stars within a distance of $0.04{\rm ~pc}$ from the MBH (denoted as GC S-stars), exclusively B-dwarfs, are spatially isotropically distributed and their orbital eccentricities follow a distribution of $f_e(e)\propto e^{2.6}$ \citep[e.g.,][]{Ghezetal08, Gillessenetal09}. The existence of these young stars is quite puzzling as star formation in the vicinity of an MBH is thought to be strongly suppressed due to the tidal force from the MBH \citep[i.e., the paradox of youth; see][]{Ghezetal08,Paumard06}. It is of great importance to address not only the formation of these stars but also the origin of their kinematics, which should encode fruitful information of the dynamical interplays between the central MBH and its environment. Young stars in the CWS (or CCWS) disk are probably formed in a previously existing massive gaseous disk due to instabilities and fragmentation developed in it \citep[e.g.,][]{Levin07,Nayakshin06,Alx08,Bonnell08}. Young binary stars in the disk(s) may migrate or be scattered into the vicinity of the central MBH \citep[e.g.,][]{MLH09} and then be tidally broken up \citep[e.g., ][]{Hills88,YT03}. One component of a broken-up binary may be ejected out as a hypervelocity star (HVS) as discovered in the Galactic halo \citep[e.g.,][]{Brown05,Edelmann05,Hirsch05}, and the other component may be captured onto a tighter orbit similar to that of the GC S-stars as proposed by \citet{Gould03}.\footnote{ Some other scenarios were also proposed to explain the orbital configuration of the GC S-stars, for example, dynamical interactions of these stars with an intermediate-mass BH in the vicinity of the central MBH (see \citealt{Merritt09,Gualandris09}) or migration of stellar binaries from the outer stellar disk to the inner region and consequent supernova explosions (see \citealt{BCL11}).} If HVSs were initially originated from a stellar structure like the CWS disk, they may be spatially located close to the disk plane \citep{Luetal10}. The current observations do show such a spatial correlation between the HVSs and the CWS disk, which suggests that majority of the HVSs originate from the CWS disk \citep{Luetal10, Zhang10}. The HVSs discovered in the Galactic halo and the GC S-stars in the vicinity of the central MBH may naturally link to each other as they may both be the products of the tidal breakup of stellar binaries in the vicinity of the central MBH \citep[e.g.,][]{GL06}. Therefore, it is interesting to simultaneously investigate the properties of the HVSs in the Galactic halo (or the GC S-stars) and their captured (or ejected) companions, and probability distribution of these properties. Under the assumption that both the HVSs and GC S-stars are the products of tidal breakup of stellar binaries, the working hypothesis in this paper, we construct a number of Monte Carlo models to simulate the tidal breakup processes of stellar binaries in the GC and check whether these models can accommodate the current observations, and make further predictions on both the companions of HVSs and that of GC S-stars for future observations.\footnote{In principle, each HVS should have a companion left in the GC and each S-star should have a companion ejected to the Galactic halo. However, these companions could have left the main sequence because of the limited lifetime and cannot be detected at the present time; and the captured companion of an HVS may even has been tidally disrupted by the central MBH and does not exist now. Considering of those cases, hereafter, the term ``companions'' may have a broad meaning in that it includes the companions of previously existed HVSs or GC S-stars as well as those detectable at the present time; and the companions of HVSs and GC S-stars may have different numbers at the present time.} This paper is organized as follows. In Section~\ref{sec:overview}, we overview the tidal breakup processes of stellar binaries in the vicinity of an MBH and the dynamical connection between the ejected and captured components. Adopting relatively realistic initial conditions, we perform a large number of three-body experiments to realize the tidal breakup processes of stellar binaries in Section~\ref{sec:num_simu}. Assuming a constant injection rate of stellar binaries into the vicinity of the central MBH and adopting the results from the three-body experiments on the ejected and captured components, we use the Monte Carlo simulations to produce both the HVSs and the GC S-stars. In Section~\ref{sec:orb_ev}, we follow the orbital evolution of the captured stars to the present time by adopting the autoregressive moving average (ARMA) model \citep{MHL10}, in which both the non-resonant relaxation (NR) and the resonant relaxation (RR) are included. The simulated GC S-stars appear to be compatible with the observations of the GC S-stars. In Section~\ref{sec:HVSs}, we investigate the effects of different binary injection models on the number ratio of the simulated HVSs to GC S-stars. The number ratio given by observations can be reproduced if the initial mass function (IMF) of the primary components of stellar binaries is somewhat top-heavy. By calibrating the injection models with observations, we estimate the number of the captured (or ejected unbound) stars, as the companions of HVSs (or GC S-stars), that could be detected in the future. We also estimate the probability to have less massive stars captured on an orbit within that of S2 in Section~\ref{sec:innermost}. Conclusions are given in Section~\ref{sec:conclusion}. For clarity, some notations of the variables that are frequently used in this paper are summarized in Table~\ref{tab:t1}. Given a physical variable $X$ (e.g., mass, velocity, semimajor axis, eccentricity), the distribution function of $X$ is denoted by $f_X (X)$ so that $f_X(X)dX$ represents the number of relevant objects with variable $X$ being in the range $X\rightarrow X+dX$. \begin{deluxetable*}{ll} \tablewidth{18.5cm} \tablecaption{Notation of Some Symbols} \tablehead{ \colhead{Symbol} & \colhead{Description} } \startdata \tabletypesize{\tiny} $M_{\bullet}$ & Mass of the central MBH \\ $m_{\rm p}$ & Mass of the primary component of an injecting stellar binary \\ $m_{\rm s}$ & Mass of the secondary component of an injecting stellar binary \\ $m$ & Total mass of an injecting stellar binary, i.e., $m_{\rm p}+m_{\rm s}$ \\ $R$ & $m_{\rm s}/m_{\rm p}$ \\ $a_{\rm b,ini}$ & Initial semimajor axis of an injecting stellar binary \\ $r_{\rm p,ini}$ & Initial pericenter distance of the mass center of the injecting stellar binary to the MBH \\ $a_{\rm b-\bullet,ini}$ & Initial semimajor axis of the orbit of an injecting stellar binary rotating around a central MBH \\ $v_{\rm \infty,ini}$ & Initial velocity of the injecting stellar binary at infinity if the binary is on a hyperbolic orbit \\ $E_{\rm ini} $ & Initial energy of the stellar binary \\ $r_{\rm tb}$ & Tidal radius for the stellar binary \\ $D$ & Orbital penetration parameter of the injecting stellar binary ($\equiv 100r_{\rm p,ini}/r_{\rm tb}$)\\ $\alpha$ & Exponent of the power-law distribution of $a_{\rm b,ini}$ \\ $\beta$ & Exponent of the power-law distribution of $r_{\rm p,ini}$ \\ $\gamma$ & Exponent of the power-law distribution of $m_{\rm p}$ \\ & \\ \hline $m_{\rm g} $ & Mass of the component that gains energy during the tidal breakup of a stellar binary\\ $m_{\rm l} $ & Mass of the component that loses energy during the tidal breakup of a stellar binary\\ $q$ & $m_{\rm l}/m_{\rm g}$ \\ $\delta E$ & Exchange energy between the two components during the tidal breakup of a stellar binary \\ & \\ \hline $m_{\rm ej}$ & Mass of the ejected star after the tidal breakup of a stellar binary \\ $m_{{\rm cap}}$ & Mass of the captured star after the tidal breakup of a stellar binary \\ $v_{\infty}$ & Velocity of the ejected component at infinity \\ $a_{{\rm cap}} $ & Orbital semimajor axis of the captured component \\ $a_{\rm cap,0}$ & Orbital semimajor axis of the captured component if the injecting binary is initially on a parabolic orbit \\ $e_{{\rm cap}} $ & Orbital eccentricity of the captured component \\ $N_{{\rm HVS}}^{{\rm tot}}$ & Simulated total number of the ejected stars given a mass range\\ $N_{{\rm cap}}^{{\rm tot}}$ & Simulated total number of the captured stars given a mass range\\ $N_{{\rm HVS}}^{\rm obs}$ & Simulated number of the detectable HVSs at the present time for given selection criteria\\ $N_{{\rm cap}}^{\rm obs}$ & Simulated number of the detectable captured stars at the present time for given selection criteria \\ $F_{{\rm HVS}}^{{\rm lt}}$ & Simulated fraction of the ejected stars that survive to the present time on the main sequence \\ $F_{{\rm cap}}^{{\rm lt}}$ & Simulated fraction of the captured stars that survive to the present time on the main sequence \\ $F_{{\rm cap}}^{{\rm td}}$ & Simulated fraction of the captured stars that have already been tidally disrupted until the present time \\ $F_{{\rm cap}}^{\rm obs}$ & Simulated fraction of the captured stars that can be detected at the present time for given selection criteria\\ \enddata \label{tab:t1} \end{deluxetable*} \section{Overview: tidal breakup of stellar binaries in the vicinity of an MBH} \label{sec:overview} A stellar binary may be broken up if it approaches an MBH within a distance of $r_{\rm tb}=a_{\rm b}(3M_\bullet/m)^{1/3}$, where $M_\bullet$ is the mass of the MBH, $a_{\rm b}$ is the semimajor axis of the binary, $m=m_{\rm g}+m_{\rm l}$ is the total mass of the binary, and $m_{\rm g}$ and $m_{\rm l}$ are the masses of the two components of the binary, respectively. During the breakup, one component of the binary, denoted as $m_{\rm g}$ here, gains energy, and the other component $m_{\rm l}$ loses energy. For an injecting stellar binary that is initially on a parabolic orbit relative to the MBH, the velocity of the binary mass center at its periapsis to the MBH ($\sim r_{\rm tb}$) is $v_{\rm tb}\sim (GM_\bullet/r_{\rm tb})^{1/2}$. The component $m_{\rm g}$ receives a velocity change on the order of $\delta v_{\rm g} \sim (m_{\rm l}/m)\sqrt{Gm/ a_{\rm b}}$ if the eccentricity of the stellar binary is $0$, and it gains energy $\delta E\sim m_{\rm g} v_{\rm tb}\delta v_{\rm g}$. The other component $m_{\rm l}$ loses the same amount of energy $\delta E$. If $\delta E$ is sufficiently large, the component $m_{\rm g}$ may manifest itself as an HVS with velocity at infinity $v_{\infty}\sim \sqrt{2 \delta E/m_{\rm g}}$ if ignoring the deceleration due to the Galactic gravitational potential. The root mean square (rms) of $v_{\infty}$ is approximately \begin{eqnarray} \left<v^2_{\infty}\right>^{1/2} & \sim & v_{\infty,0} \left(\frac{0.1{\rm AU}}{a_{\rm b}}\right)^{1/2} \left(\frac{m}{6M_{\odot}}\right)^{1/3}\nonumber\\ &\times &\left(\frac{2m_{\rm l}}{m}\right)^{1/2} \left(\frac{M_\bullet}{4\times 10^6M_{\odot}}\right)^{1/6} g(D), \label{eq:vhvs} \end{eqnarray} where $ v_{\infty,0}=2596{\rm km\,s^{-1}}$ and $g(D)$ is given by \citet[][]{Bromley06} for injecting binaries on hyperbolic orbits with initial velocities at infinity of $250{\rm km\,s^{-1}}$, i.e., \begin{eqnarray} g(D)& = & 0.774+0.0245D-8.99\times 10^{-4}D^2 \nonumber \\ & & +1.32\times 10^{-5}D^3-8.82\times10^{-8}D^4 \nonumber \\ & & +2.15\times 10^{-10}D^5, \label{eq:fd1} \end{eqnarray} where the penetration parameter $D\equiv 100r_{\rm p,ini}/r_{\rm tb}$ characterizes the minimum distance where the binary approaches the MBH, and $r_{\rm p,ini}$ is the initial pericenter distance of the binary. The rms velocity $\left<v^2_{\infty}\right>^{1/2}$ apparently depends on the semimajor axis, the total mass and the mass ratio of the stellar binary, and the penetration parameter $D$. The exact value of $v_{\infty}$ of the ejected component for any given stellar binary also depends on the relative orientation of the stellar binary orbital plane to the orbital plane of the binary rotating around the MBH and the orbital phases of the two components at the time of its breakup. This dependence introduces a scatter of $v_{\infty}$ around the value $\left< v^2_{\infty} \right>^{1/2}$ given by Equation (\ref{eq:vhvs}), as the orbital orientations of the injecting stellar binaries are probably random and the orbital phases of the two components are not fixed at the breakup time. Numerical simulations have shown that this scatter is approximately Gaussian with a dispersion of $\sigma_{v_{\infty}}\sim 0.2 \left<v^2_{\infty} \right>^{1/2}$ \citep{Bromley06,Zhang10}, where the binary orbital orientations are assumed to be randomly distributed. The symmetry of the orbital phases of the two binary components (always at the opposite side to the mass center of the binary) ensures the same probability of receiving energy for each star, which leads to the same ejection probability for both components if the injecting binaries are initially on parabolic orbits \citep{Sari10,Sari12}. Stellar binaries on orbits bound to the MBH may experience multiple close encounters with the MBH and the binary semimajor axes and eccentricities may be cumulatively excited to larger values until finally being broken up \citep{Zhang10}. The distribution of $v_{\infty}$ for the ejected stars, produced during the first encounters of the binaries with the MBH, follows a fitting formula similar to Equation (\ref{eq:fd1}) over $D\sim 20$--$150$, i.e., $g(D) \propto 1-(D/256)^2$, as the initial bounding energy of the injecting stellar binaries is still significant~\citep[for details, see][]{Zhang10}. For multiple encounters, the energy exchange $\delta E$ between the two components is determined by the properties of the stellar binaries at the final revolutions. Our simulations show that $\left<v^2_{\infty}\right>^{1/2}$ of those ejected components for stellar binaries broken up within $1000$ revolutions around the MBH still follows Equation (\ref{eq:vhvs}), but $g(D)$ is now best fitted by \begin{eqnarray} g(D)& = &0.912-2.41\times10^{-4} D -4.49\times10^{-5}D^2\nonumber \\ & &+2.68\times10^{-7} D^3-4.42\times10^{-10} D^4, \label{eq:fd2} \end{eqnarray} for $D<300$. For stellar binaries on bound orbits, the light component has a larger probability to escape away from the MBH because the specific energy it could gain is generally larger than that of the heavy component in a counterpart case (see Equation~\ref{eq:vhvs}). However, the difference in the ejection probability for the two components of the stellar binaries is significant only when the mass ratio of the massive ones to the light ones $\mathrel{\mathpalette\oversim>} 5$ and $2\delta E/m_{\rm g}$ is close to its initial bounding energy $GM_{\bullet}/2a_{\rm b-\bullet,{\rm ini}}$, where $a_{\rm b-\bullet,{\rm ini}}$ denotes the initial semimajor axis of the binary system composed of a stellar binary and the MBH \citep[see also][]{Antonini11, Sari12}. The component $m_{\rm l}$ of a broken-up stellar binary loses energy by an amount of $\delta E$ and it is captured onto a tighter orbit with semimajor axis $a_{\rm cap}$. According to the energy conservation law, we roughly have \begin{equation} \frac{1}{2}m_{\rm g} v_{\infty}^2-\frac{Gm\lM_{\bullet}}{2a_{\rm cap}} \simeq E_{{\rm ini}}, \label{eq:dk_b} \end{equation} where $E_{{\rm ini}}$ is the initial energy of the stellar binary, and it is $\sim \frac{1}{2}m v_{\infty,{\rm ini}}^2$ if the binary is initially on a hyperbolic orbit, or $\sim -GmM_{\bullet}/(2a_{\rm b-\bullet,{\rm ini}})$ if on a bound orbit. The initial internal mechanical energy of the stellar binary $-\frac{Gm_lm_g}{2a_{\rm b,ini}}$ is ignored in Equation (\ref{eq:dk_b}). We now have the general form for $v_{\infty}$ as $v_{\infty} \sim \sqrt{2(\delta E+\frac{m_{\rm g}}{m} E_{{\rm ini}})/m_{\rm g}}$. If $|E_{{\rm ini}}|\ll \delta E$, the semimajor axis of the captured star is \begin{eqnarray} a_{\rm cap}& \simeq & a_{\rm cap,0} = q\frac{GM_{\bullet}}{v^2_\infty} \nonumber \\ & = & 3500q {\rm AU} \left(\frac{M_{\bullet}}{4\times 10^6M_{\odot}}\right) \left(\frac{1000{\rm km\,s^{-1}}}{v_{\infty}}\right)^2, \label{eq:acap0} \end{eqnarray} where $q\equiv m_{\rm l}/m_{\rm g}$. For the cases considered in this paper, the injecting stellar binaries are either initially on hyperbolic orbits (but close to parabolic ones) or from stellar structures like the CWS disk, and thus $a_{\rm cap}\sim a_{\rm cap,0}$ as approximately $|E_{{\rm ini}}|\ll \delta E$. Equation (\ref{eq:acap0}) shows the connection between the properties of the captured stars left in the GC and that of their ejected companions in the Galactic bulge and halo. For those HVSs discovered in the Galactic halo with $v_{\infty} \sim 700$-$1000{\rm km\,s^{-1}}$,\footnote{ The estimated $v_{\infty}$ for those detected HVSs depends on the Galactic potential model adopted, the values here are obtained from the Galactic potential model given by \citet{Xue08}.} their companions left in the GC may be initially on orbits with semimajor axis in the range of $\sim 3500$-$7000{\rm AU}$ as stellar binaries with extreme mass ratios are rare. For the innermost S-star, i.e., the S2, of which the semimajor axis is $\sim 1000{\rm AU}$, its companion ejected out should have $v_{\infty} \sim 1900{\rm km\,s^{-1}}$ if $q \sim1$, and $\sim 600{\rm km\,s^{-1}}$ if $q \sim 0.1$, respectively. For stellar binaries initially tightly bound to the MBH, we may have $\delta E + \frac{m_{\rm g}}{m} E_{{\rm ini}}\mathrel{\mathpalette\oversim<} 0$, the component gaining energy either remains bound to the MBH or is ejected out with low velocities. The distribution of HVS properties is directly connected to the distribution of S-star properties (note that here we do not mean that an observed HVS in the Galactic halo is directly associated with an observed GC S-star as the products of the tidal breakup of the same binary star). According to Equation (\ref{eq:acap0}), the distribution of the semimajor axis of the captured stars $f_{a_{\rm cap}}$ is related to the distribution of the velocity of HVSs at infinity $f_{v_{\infty}}$ if $m_{\rm g}\sim m_{\rm l}$, i.e., \begin{equation} f_{v_{\infty}}(v_{\infty})\propto \left. a_{\rm cap}^{3/2}f_{a_{\rm cap}}(a_{\rm cap})\right|_{a_{\rm cap}=\frac{GM_{\bullet}}{v_\infty^2}}, \label{eq:va} \end{equation} which suggests that any one of the two distributions above can be inferred from the other one. The velocity distribution of the ejected stars are mainly determined by the initial sets on the distributions of $a_{\rm b,ini}$ and $r_{\rm p,ini}$ since \begin{equation} \left<v^2_{\infty}\right>^{1/2}\propto a_{\rm b,ini}^{-1/2}g(D), \end{equation} where $g(D)$ denotes the dependence of the rms velocity $\left<v^2_{\infty} \right>^{1/2}$ on the penetration parameter $D$, as shown in Equations (\ref{eq:fd1}) and (\ref{eq:fd2}) for the cases of injecting binaries initially on hyperbolic orbits but close to parabolic ones and bound orbits like the stars in the CWS disk, respectively. The fitting forms of $g(D)$ obtained from numerical experiments (Equations (\ref{eq:fd1}) and (\ref{eq:fd2})) are decreasing functions in the range of $20<D<150$ or $20<D<300$, and thus $g(D)$ may be approximated as a monotonically decreasing function. We assume that the initial distribution of $a_{\rm b,ini}$ and $r_{\rm p,ini}$ are $f_{a_{\rm b}}(a_{\rm b,ini}) \propto a_{\rm b,ini}^\alpha$ and $f_{r_{\rm p}}(r_{\rm p,ini}) \propto r_{\rm p,ini}^\beta$, respectively; and the probability of a stellar binary with semimajor axis $a_{\rm b,ini}$ broken up by the central MBH at a penetration distance $D$ is only a function of $D$, i.e., $f_D(D)$ \citep[see][]{Bromley06}. If $q\sim 1$ and ignoring the scatter of $v_{\infty}$ around $\left<v_{\infty}^2\right>^{1/2}$ (i.e., $v_{\infty}\sim \left<v^2_{\infty}\right>^{1/2}$), the velocity distribution of the ejected components can be obtained as \begin{eqnarray} f_{v_{\infty}}(v_{\infty}) & \propto & \frac{\partial}{\partial v_\infty}\int\int f_{a_{\rm b}}(a_{\rm b,ini})f_{r_{\rm p}}(r_{\rm p,ini}) \times \nonumber \\ & & f_D(D) da_{\rm b,ini} dr_{\rm p,ini} \nonumber \\ & \propto & \frac{\partial}{\partial v_\infty} \int a_{\rm b,ini}^{\alpha+\beta+1} da_{\rm b,ini} \int D^{\beta} f_D(D) dD \nonumber \\ & \propto &v_\infty^{-2\alpha-2\beta-5}, \label{eq:fV} \end{eqnarray} and this relation is valid only if $g(D)$ is a monotonically decreasing function and it is independent of the detailed form of $f_D(D)$. Similarly, we also have \begin{equation} f_{a_{\rm cap}}(a_{\rm cap}) \propto a_{\rm cap}^{\alpha+\beta+1}, \label{eq:fac} \end{equation} which is consistent with the simple relation given by Equation (\ref{eq:va}). The estimated slope of $f_{v_{\infty}}(v_{\infty})$ (or $f_{a_{\rm cap}}(a_{\rm cap})$) above is not affected by taking account of the Gaussian-like scatter of $v_{\infty}$ around $\left<v^2_{\infty}\right>^{1/2}$ as the distribution is a power law. If considering of the various mass ratios among the injecting binaries (see Equations (\ref{eq:vhvs}) and (\ref{eq:acap0})), however, the resulted slope of $f_{v_{\infty}}(v_{\infty})$ may be somewhat flatter than the simple estimates above. Note also that a larger $\beta$ may correspond to a slower migration/diffusion of stellar binaries into the low angular momentum orbits or the vicinity of the central MBH, and lead to fewer HVSs at the high-velocity end and fewer captured stars in smaller distances to the MBH. The periapsis of a captured star $m_{\rm l}$ is roughly $\sim r_{\rm tb}$ and the characteristic eccentricity of the captured star is \begin{equation} \bar{e}_{{\rm cap}}\sim 1-\frac{r_{\rm tb}}{a_{\rm cap}}\simeq 1- \frac{2.8}{q^{1/3}(1+q)^{2/3}} \left(\frac{m_{\rm l}}{M_{\bullet}} \right)^{1/3}. \label{eq:ecc} \end{equation} the $\bar{e}_{{\rm cap}}$ depends on $q$ and the mass ratio of the captured star to the MBH. Considering of the Gaussian-like scatter in $v_{\infty}$ and correspondingly the scatter in $a_{\rm cap}$, the probability that the breakup of a stellar binary with given semimajor axis and mass of each component results in a captured star with eccentricity $<e_{\rm cap}$ is roughly \begin{eqnarray} P(<e_{\rm cap})=\frac{1}{2}{\rm erfc}\left[\frac{\sqrt{(1-e_{\rm cap})/ (1-\bar{e}_{{\rm cap}})}-1}{\sqrt{2}\sigma_{v_{\infty}}/ \langle v^2_{\infty}\rangle^{1/2}}\right], \label{eq:ecprob} \end{eqnarray} where $\bar{e}_{{\rm cap}}$ is given by Equation (\ref{eq:ecc}) and $\sigma_{v_{\infty}}/ \langle v^2_{\infty}\rangle^{1/2}\simeq 0.2$. To capture an S2-like star (i.e., $e_{\rm cap}\simeq 0.887$ and $m_{\rm l}\sim 15M_{\odot}$; see \citealt{Ghezetal08}; \citealt{Gillessenetal09}) directly through the tidal breakup of stellar binaries, it is necessary to have $m_{\rm g}\gg m_{\rm l} \sim 15 M_{\odot}$ ($q\ll 1$) and the probability is $\sim 0.25$ if $q=0.1$ according to Equation (\ref{eq:ecprob}) \citep[see][]{Gould03}. The probability to capture stars onto orbits with $e_{\rm cap}<0.8$ is only $\sim 10^{-17}$ if $m_{\rm g}=m_{\rm l}=15M_{\odot}$; and $\sim6 \times 10^{-3}$ even if $m_{\rm g}=10m_{\rm l}=150M_{\odot}$ (see also \citealt{Gould03}). Since the number of GC S-stars is only on the order of a few tens, it is difficult to produce all the nine observed GC S-stars with eccentricities $<0.8$ \citep{Gillessenetal09} directly by the tidal breakup of stellar binaries. In addition, the captured stars may initially remain on a disk plane if their progenitor binaries are originated from disk-like stellar structure(s) as suggested by \citet{Luetal10}, which is different from the isotropic distribution of the GC S-stars. Therefore, additional physical mechanism is required to further make the captured stars evolve to orbits with lower eccentricities and spatially isotropically distributed if the GC S-stars are really originated from the tidal breakup of binary stars. The processes, initially proposed by \citet{RT96}, may cause the captured stars dynamically evolving to their present orbits as discussed by a number of authors \citep{Levin07,HA06,KT11}. In Section~\ref{sec:orb_ev}, we will take into account the relaxation processes, including RR, to approximately follow the dynamical evolution of each ``GC S-star'' after its capture due to the tidal breakup of stellar binaries; we then check whether the eccentricity and spatial distributions of those surviving ``GC S-stars'' are compatible with current observations. Note here we use the quotes around the term GC S-stars to represent all of those captured stars with mass in the range of $\sim 7$-$15M_{\odot}$; while the simulated GC S-stars (without quotes) represent those with mass $\sim 7$-$15M_{\odot}$ surviving to the present time, which presumably correspond to the observed ones (see Section~\ref{sec:orb_ev}). \section{Monte Carlo Simulations}\label{sec:num_simu} In this section, we first adopt Monte Carlo simulations to realize the tidal breakup processes and generate HVSs and ``GC S-stars'', and then we investigate in detail the connection between the simulated HVSs and ``GC S-stars''. We use the code DORPI5 based on the explicit fifth (fourth)-order Runge--Kutta method \citep{DP80,Hairer93} to calculate the three-body interactions between a stellar binary and the central MBH. For details of the numerical calculations, see \citet{Zhang10}. The successive dynamical evolution of the captured stars in the GC and the kinematic motion of the produced HVSs in the Galactic potential will be discussed in Sections~\ref{sec:orb_ev} and \ref{sec:HVSs}, respectively. \subsection{Initial Settings}\label{subsec: init_set} The mass of the central MBH is set to be $4\times10^6M_{\odot}$ throughout the numerical calculations in this paper \citep{Ghezetal08,Gillessenetal09}. For the injecting stellar binaries, the initial conditions are set as follows: \begin{itemize} \item The distribution of the semimajor axes $a_{\rm b,{\rm ini}}$ follows the \"{O}pik law, i.e., $\alpha=-1$~\citep[e.g.,][]{KF07}. \item The mass distribution of the primary stars $m_{\rm p}$ follows a power law function, $f_{m_{\rm p}}(m_{\rm p}) \propto m_{\rm p}^{\gamma}$. The distribution of the secondary star ($m_{\rm s}$) or the mass ratio $R\equiv m_{\rm s}/m_{\rm p}$ can be described by two populations: (1) a twin population, i.e., about 40\% of binary stars have $R\sim 1$, and (2) the rest binaries, which follow a distribution of $f_R(R)\sim {\rm constant}$~\citep{KF07,Kiminki08,Kiminki09}. \item The initial eccentricity of the injecting binary is assumed to be $e_{{\rm ini}}=0$, as adopted in previous works (e.g., \citealt{Bromley06}; \citealt{Antonini10}).\footnote{Alternatively assuming the initial eccentricities $\sim 0.3$, the velocities of the resulted HVSs from the four models are roughly smaller than those obtained for $e_{{\rm ini}}\sim 0$ by $\mathrel{\mathpalette\oversim<} 10\%$.} \item The orientation of the inner binary orbital plane is chosen to be uniformly distributed in $\cos\phi$ for $\phi\in (0,\pi)$. \end{itemize} For the orbits of the injecting stellar binaries relative to the central MBH, the initial conditions are set as follows: \begin{itemize} \item The stellar binaries are assumed to be initially injected from either disk-like stellar structures (similar to the CWS disk) or infinity. If they were from structures like the CWS disk, the semimajor axes $a_{\rm b-\bullet,{\rm ini}}$ follows a power-law distribution proportional to $a_{\rm b-\bullet,ini}^{-2.3}$ in the range of $\sim 0.04$-$0.5{\rm ~pc}$ according to current observations on the CWS disk \citep{LuJ09,Bartko09}. If they were from infinity, i.e., unbound to the MBH, their initial velocities at infinity are set to $v_{\rm \infty,ini}=250{\rm km\,s^{-1}}$. \item If the injecting binaries were from disk-like stellar structures, the orientations of their orbits relative to the MBH are assumed to satisfy a Gaussian distribution around the central planes of the stellar disks with a standard deviation of $12{^{\circ}}$ \citep[cf.][]{LuJ09,Bartko09}. The planes of the host disks are assumed to be the same as the two planes that best fit the observations, i.e., $(l,b) = (311{^{\circ}}, -14{^{\circ}})$ and $(176{^{\circ}}, -53{^{\circ}})$, respectively, and these two planes are consistent with the CWS disk plane and the plane of the northern arm (Narm) of the mini-spiral in the GC (or the outer warped part of the CWS disk; \citealt{Luetal10, Zhang10}).\footnote{Note that \citet{Brown12} recently reported five new unbound HVSs discovered in the Galactic halo and re-analyzed the HVSs previously discovered. According to this new study, there are $17$ unbound HVSs in the northern sky and they are still consistent with being located on two planes revealed by \citet{Luetal10}. That is, one of the disk planes is consistent with the CWS disk plane, while the other disk plane is more consistent with the warped outer part of the CWS disk and slightly deviates from the Narm plane. In this paper, we do not distinguish the Narm plane from the warped outer part of the CWS disk.} The injection rates from these two disks are assumed to be the same. \item The periapsis that the injecting stellar binaries approach the MBH is simply assumed to follow a power law distribution, $f_{r_{\rm p}}(r_{\rm p,ini})\proptor_{\rm p,ini}^{\beta}$ and $\beta>0$. A larger value of $\beta$ corresponds to a smaller fraction of the injecting stellar binaries that could approach the immediate vicinity of the central MBH. It is still not clear which mechanism is responsible for the migration (or diffusion) of stellar binaries into the vicinity of the central MBH, although the secular instability developed in a stellar disk is proposed to be a viable one \citep{MLH09}. Instead of incorporating the detailed migration/diffusion process of the stellar binaries in the Monte Carlo simulations below, we choose to parameterize the migration/diffusion process qualitatively by different values of $\beta$ and a larger $\beta$ corresponds to a slower migration/diffusion process. \end{itemize} \begin{deluxetable}{lcccc} \tablecaption{Different Injection Models for Tidal Breakup of Binaries} \tablehead{ \colhead{Model} & \colhead{$\gamma$} & \colhead{$\beta$} & \colhead{$a_{\rm b-\bullet,ini}({\rm ~pc})$} & \colhead{$v_{\rm \infty,ini}({\rm km\,s^{-1}})$\tablenotemark{a}}} \startdata Unbd-MS0 & -2.7 & 0 & $\cdots$ & 250 \\ Disk-MS0 & -2.7 & 0 & 0.04-0.5 & $\cdots$ \\ Disk-TH0 & -0.45 & 0 & 0.04-0.5 & $\cdots$ \\ Disk-TH2 & -0.45 & 2 & 0.04-0.5 & $\cdots$ \enddata \tablenotetext{a}{For the Unbd-MS0 model, the injecting stellar binaries have initial velocities of $250{\rm km\,s^{-1}}$ at infinity.} \label{tab:t2} \end{deluxetable} In this section, we perform Monte Carlo simulations by adopting four sets of initial conditions (as listed in Table~\ref{tab:t2}). In the first model, the stellar binaries are assumed to be injected from infinity with initial velocity of $v_{\rm \infty,ini} =250{\rm km\,s^{-1}}$. For the primary components of the injecting binaries, we adopt the Miller Scalo IMF (e.g., \citealt{Kroupa}). This model is denoted as ``Unbd-MS0''. For the other three models, the stellar binaries are assumed to be originated from stellar structures like the CWS disk, and the IMF of the primary components is either set to be the Miller Scalo IMF or a top-heavy IMF with a slope of $\gamma=-0.45$ as suggested by recent observations of the disk stars \citep[see][]{Bartko10}. The slope of the initial distribution of the pericenter distance $\beta$ is set to be either $0$ or $2$. These models are denoted as ``Disk-MS0'', ``Disk-TH0'', and ``Disk-TH2'', respectively. The total number of three-body experiments is $10^5$ for each model with the initial settings described above. By comparing the results obtained from those different models, one may be able to distinguish the effects of different settings on the IMF and the injection of stellar binaries. If not specified, those ejected or captured stars with mass in the range of $\sim 3$-$15M_{\odot}$ are recorded, thus both the HVSs with mass $\sim 3$-$4M_{\odot}$ and the captured stars with mass $\sim 7$-$15M_{\odot}$, corresponding to the currently detected ones, can be taken into account simultaneously. The ejected or captured stars with mass in the range of $\sim 4$-$7M_{\odot}$ are also considered for completeness. For other ejected or captured stars with mass out of the range of $3$-$15M_{\odot}$, they may be either too faint to be detected or too massive with too short lifetime and thus with too small probability to survive. \subsection{Numerical Results}\label{subsec:results} \begin{figure*} \centering \includegraphics[scale=0.5]{f1.eps} \caption{ Number distributions of broken-up stellar binaries in the $v_{\infty}$-$a_{\rm cap}$ plane, where $v_{\infty}$ is the velocity at infinity of the ejected component and $a_{\rm cap}$ is the semimajor axes of its captured companion. Panels (a)--(d) show results obtained from the Unbd-MS0, Disk-MS0, Disk-TH0, and Disk-TH2 models, respectively. The total number of the three-body experiments is $10^5$ for each model. The solid red line in each panel shows the estimation according to Equation (\ref{eq:dk_b}) for $m_{\rm g}=m_{\rm l}$ by assuming $v_{\rm \infty,ini} = 250{\rm km\,s^{-1}}$ in panel (a), and $a_{\rm b-\bullet,{\rm ini}}=0.2{\rm ~pc}$ in panels (b)--(d), respectively. The dashed red lines above or below the solid lines represent the estimations for $m_{\rm g}/m_{\rm l}=1/2$ and $2$, respectively. The magenta lines with triangles indicate the rms of $v_{\infty}$ for each bin of $a_{\rm cap}$. The number of stars is counted in each of the $a_{\rm cap}$ and $v_{\infty}$ bins (totally $25\times25$ bins with bin size $400{\rm AU}\times 100{\rm km\,s^{-1}}$) and represented by the color brightness scales shown in the label. } \label{fig:f1} \end{figure*} \begin{figure*} \centering \includegraphics[scale=0.5]{f2.eps} \caption{ Number distributions of the captured stars in the $a_{\rm cap}$ vs. $\log(1-e_{\rm cap})$ plane, where $a_{\rm cap}$ and $e_{\rm cap}$ are the semimajor axes and eccentricities of the captured stars achieved right after their capture, respectively. Panels (a)--(d) show the results from the Unbd-MS0, Disk-MS0, Disk-TH0, and Disk-TH2 model, respectively. The red open circles represent the observed GC S-stars with $a_{\rm cap}$ smaller than $4000{\rm AU}$ at the present time \citep{Gillessenetal09}; and the magenta dashed and dotted lines are for the mean eccentricities given by Equation (\ref{eq:ecc}) for $(m_{\rm l}, q)=(10M_{\odot}, 1)$ and $(10M_{\odot}, 1/10)$, respectively. The number of stars is counted in each of the $a_{\rm cap}$ and $\log(1-e_{\rm cap})$ bins (totally $25\times25$ bins with bin size $200{\rm AU}\times 0.08$) and represented by the color brightness scales shown in the label. } \label{fig:f2} \end{figure*} Figure~\ref{fig:f1} shows the distribution of the tidally broken-up stellar binaries in the $v_{\infty}$-$a_{\rm cap}$ plane, where $v_{\infty}$ is the velocity at infinity of the ejected component and $a_{\rm cap}$ is the semimajor axis of the captured component. As shown in panel (a), the majority of the simulated $v_{\infty}$-$a_{\rm cap}$ pairs obtained from the Unbd-MS0 model are close to the one estimated from Equation (\ref{eq:acap0}) by setting $m_{\rm l}/m_{\rm g}=1$ (solid line). The main reasons for this are: (1) the majority ($70\%$) of the injecting stellar binaries have mass ratios $q=m_{\rm l}/m_{\rm g}$ in the range of (1/2, 2) under the assumption of two populations set for the stellar binaries; and (2) all the injecting binaries have the same but negligible initial energy $E_{{\rm ini}}$. A small number of $v_{\infty}$-$a_{\rm cap}$ pairs, which apparently deviate significantly away from the solid line (for $q=1$ obtained from Equation \ref{eq:acap0}), are due to the breakup of the binaries with $q$ substantially larger or smaller than $1$ (below or above the solid line). For the other three models, the simulation results do not deviate far away from the simple predictions by Equation (\ref{eq:dk_b}) (for $q=1$), except that fewer ejected stars at the high-velocity end are produced in the Disk-TH2 model than in the other models simply because not many stellar binaries can closely approach the MBH. The scatters of $v_{\infty}$ around that predicted by Equation (\ref{eq:dk_b}) in panels (b)-(d) are more significant compared with that in panel (a), which is caused by one or the combination of the effects as follows: (1) a distribution of the negative initial energy of the injecting stellar binaries originated from stellar structure like the CWS disk (panels (b)-(d)); (2) relatively more progenitor binaries have $q$ substantially larger or smaller than $1$ in the cases with a top-heavy IMF (panels (c) and (d)); and (3) fewer stellar binaries approach the immediate vicinity of the central MBH in the case of a large $\beta$ (panel (d)). As seen from Figure~\ref{fig:f1}, if the observed GC S-stars, with semimajor axes $\sim 1000$-$4000{\rm AU}$, are produced by the tidal breakup of stellar binaries, their ejected companions are expected to have $v_{\infty} \sim 1000$-$1600{\rm km\,s^{-1}}$ in the Unbd-MS0 model and $\sim 200$-$1500{\rm km\,s^{-1}}$ in the other models. The Disk-TH2 model produces fewer ejected stars with $v_{\infty}$ substantially larger than $1000{\rm km\,s^{-1}}$ compared with other models. The captured companions of the detected HVSs in the Galactic halo are more likely to have $a_{\rm cap}\sim 3000$-$8000{\rm AU}$ in the Unbd-MS0 model, which is consistent with the simple estimation by Equation (\ref{eq:acap0}), and have $a_{\rm cap}\sim1000$-$8000{\rm AU}$ in the other models. The travel/arrival time of the detected HVSs from the GC to its current location is on the order of $\sim100$~Myr \citep[e.g.,][]{Brown12b}, which suggests that their companions were captured $\sim 100$~Myr ago and the orbits of the captured companions may have been changed due to the dynamical interactions with its environment (see Section~\ref{sec:orb_ev}). As shown in Figure~\ref{fig:f1}, for those captured stars with $a_{\rm cap}\mathrel{\mathpalette\oversim>} 10,000{\rm AU}$, the probability that they have ejected companions with $v_{\infty} > 700$-$1000{\rm km\,s^{-1}}$ is negligible. Figure~\ref{fig:f2} shows the distribution of those captured stars obtained from each model in the $a_{\rm cap}$ versus $\log(1-e_{\rm cap})$ plane, where $a_{\rm cap}$ and $e_{\rm cap}$ are their semimajor axes and eccentricities achieved right after they were captured, respectively. Relatively more captured stars with low eccentricities are produced by the Disk-MS0 model than by the Unbd-MS0 model (see panels (a) and (b)) mainly because the injecting binaries can be broken up at relatively larger distance in the Disk-MS0 model due to multiple encounters. And relatively more captured stars with low eccentricities and $a_{\rm cap}$ are produced in the Disk-TH model than that in the Unbd-MS0 model (see panels (a) and (c)) because there are more injecting binaries with mass ratio $q$ substantially less than $1$ and $m_{\rm l} \sim 7$-$15 M_{\odot}$ for a top-heavy IMF than that for the Miller Scalo IMF. The Disk-TH2 model produces relatively more captured stars with smaller eccentricities for any given $a_{\rm cap}$ than the Disk-TH0 model (as shown in panels (c) and (d)), as those binaries are generally broken up at even larger distances in the Disk-TH2 model because fewer binaries can approach the very inner region due to the steepness of the adopted $f_{r_{\rm p}}(r_{\rm p,ini})$. However, the eccentricities of those captured stars, even produced in the Disk-TH2 model, are still statistically significantly higher than that of the observed GC S-stars. The orbits of a number of GC S-stars, including S2, can be directly produced in the Disk-TH0 model and the Disk-TH2 model if the injection rate of binaries is around a few times $10^{-5}$ to $10^{-4}{\rm yr}^{-1}$ as set for those models (see similar rates obtained by \citealt{Bromley12}). According to Figure~\ref{fig:f2}, apparently it is extremely difficult to produce ``GC S-stars'' with $e_{\rm cap}<0.8$ directly through the tidal breakup mechanism of stellar binaries in the vicinity of the MBH. Note also that fewer captured stars with $a_{\rm cap}< 1000{\rm AU}$ are produced in the Disk-TH2 model compared with those in other models because stellar binaries are harder to approach the innermost region than that in other models (see panel (d) in Figures \ref{fig:f1} and \ref{fig:f2}). According to the simulations above, we find that $\sim50\%$ of the detected HVSs should have captured companions in the GC with mass $\sim 3$-$4M_{\odot}$ as shown in the left panel of Figure~\ref{fig:f3}. And similarly $\sim 60\%$ of the observed GC S-stars should have ejected companions with mass $\sim 7$-$15M_{\odot}$ as shown in the right panel of Figure~\ref{fig:f3}. To find the possible counterparts of those current observed GC S-stars and HVSs, we will focus on the ejected stars with mass $\sim 7$-$15M_{\odot}$ in the Galactic bulge and halo and the captured stars with mass $\sim 3$-$4M_{\odot}$ in the GC. For completeness, we also count the ejected and captured stars with mass $\sim 4$-$7M_{\odot}$ produced in all the models (see Table \ref{tab:t4}). \begin{figure*} \centering \includegraphics[scale=0.5]{f3.eps} \caption{ Mass probability distribution $P(m_{{\rm cap}})$ of the captured companions of the $\sim 3$-$4M_{\odot}$ HVSs (left panel) and mass probability distribution $P(m_{{\rm ej}})$ of the ejected companions of the $\sim 7$-$15M_{\odot}$ captured stars (right panel). The solid (red), dotted (blue), dashed (magenta) and dot-dashed (cyan) lines represent the results obtained from the Unbd-MS0, the Disk-MS0, the Disk-TH0, and the Disk-TH2 models, respectively. } \label{fig:f3} \end{figure*} \section{Orbital evolution of the captured stars}\label{sec:orb_ev} The orbits of the captured stars produced by the tidal breakup of stellar binaries may evolve due to dynamical interactions with the surrounding environments. In principle, two-body interactions between stars may cause exchanges of their angular momenta and energy. However, the timescale of the two-body relaxation ($\sim 10^9$~yr) is too long for it to be effective in changing the orbits of captured stars within their main-sequence lifetime \citep[e.g.,][]{HA06,Yuetal07}. The RR is an important dynamical process naturally resulted from the coherent torques between orbital averaged mass wires of stars moving in near-Keplerian potential proposed by \citet{RT96}, which can lead to changes in both the eccentricities (scalar RR) and orientations (vector RR) of stars moving in the GC \citep[e.g.,][]{HA06}. The RR appears much more effective in changing the orbital configuration of the ``GC S-stars'' than the non-resonant two-body relaxation \citep[NR;][]{HA06,Perets09c,KT11}. Therefore, the scalar and vector RR may be crucial in the follow-up dynamical evolution of the orbits of the captured stars. The vector RR timescale is $\sim 1$-$10$~Myr in the region hosting the ``GC S-stars'', about one order of magnitude smaller than the scalar RR timescale \citep{HA06, Yuetal07}, and thus the captured stars can evolve to an isotropic distribution on a timescale of $\sim10$~Myr \citep{HA06, Perets09c, KT11} even if they were originally on a plane-like structure. In this paper, we assume that the isotropic distribution of the GC S-stars can always be reproduced through the vector RR of the simulated captured stars within a timescale shorter than their lifetime. It has been suggested that the high-eccentric orbits of captured stars can dynamically evolve to that of the observed GC S-stars through the scalar RR within $\sim20$~Myr \citep[e.g.,][]{Perets09c}. However, previous studies assume a simple distribution of the initial eccentricities and semimajor axes of the captured stars \citep[e.g.,][]{Perets09c, MHL10}. In this Section, we adopt the distribution of $e_{\rm cap}$ and $a_{\rm cap}$ resulted from the injection models studied in Section~\ref{sec:num_simu}. We follow the evolution of $e_{\rm cap}$ and $a_{\rm cap}$ by taking both the RR and NR into account, and then compare the $e_{\rm cap}$ and $a_{\rm cap}$ distributions of the captured stars surviving to the present time with that of the observed GC S-stars. We adopt the ARMA model first introduced by \citet{MHL10} to perform Monte Carlo simulations of the long-term evolution of the captured stars. In the ARMA model, the RR phase and the NR phase are unified, and the general relativistic (GR) precession of stars in the potential of the central MBH is also simultaneously included. The ARMA model is characterized by the following three parameters: (1) the autoregressive parameter $\phi_1$; (2) the moving average parameter $\theta_1$; and (3) the parameter $\sigma_1$, which is the variance of a random variable $\epsilon^{(1)}$ following the normal distribution. In the ARMA model, $\epsilon^{(1)}$ represents the random walk motion of the NR phase. At a time step of one orbital period of a star, the variation in the absolute value of its angular momentum is \begin{equation} \Delta_1 J_t=\phi_1\Delta_1J_{t-1}+\theta_1\epsilon^{(1)}_{t-1} +\epsilon^{(1)}_t, \end{equation} and \begin{equation} \phi_1=\exp{(-\frac{\delta t_{\rm P}}{S t_\phi})}, \end{equation} \begin{equation} t_{\phi}=f_{\phi}{\rm min}[t_{\rm prec}(a,e),t_{\rm prec}(a,\tilde{e})], \end{equation} \begin{equation} S=\frac{1}{1+\exp{\left[-k\left(e-e_{\rm crit}\right)\right]}},\footnote{ \rm There is a misprint in the original form of $S$ given by \citet[][see their Equation 33]{MHL10}, i.e., the minus sign before $k$ was missing. } \end{equation} \begin{equation} e_{\rm crit}(a,e)=\sqrt{\frac{\ln\Lambda}{A_{\rm NR}A_{\tau}^2}} \left(\frac{\delta t_{\rm P}}{t_{\phi}}\right), \end{equation} \begin{equation} \theta_1=-\exp\left[{-\frac{f_\theta}{2}\sqrt{\frac{1}{\phi_1^2}+\phi^2_1-2+ \frac{4(1-\phi^2_1)\tau^2\delta t_{\rm P}^2}{\sigma^2_1}}}\right], \end{equation} \begin{equation} \tau=A_{\tau}\frac{m_*}{M_{\bullet}}\frac{\sqrt{N_{<}}}{\delta t_{\rm P}}e, \end{equation} \begin{equation} \sigma_1=f_\sigma \frac{m_*}{M_\bullet}\sqrt{\frac{N_< \ln\Lambda}{A_{\rm NR}}}, \end{equation} \begin{equation} f_{\sigma}=0.52+0.62e-0.36e^2+0.21e^3-0.29\sqrt{e}, \end{equation} where $f_{\phi}=0.105$, $f_\theta=1.2$, $k=30$, $A_{\rm NR}=0.26$, $A_{\tau} =1.57$, $\Lambda=M_{\bullet}/m_*$, $m_*=10M_{\odot}$ is the averaged mass of the field stars, $a$ and $\delta t_{\rm P}$ are the semimajor axis and orbital period of the star, $\tilde{e}$ is the median value of the eccentricity of the field stars and it is $\sqrt{1/2}$ for a thermal distribution, $t_{\rm prec}$ is the combined precession timescale for the Newtonian precession and the general relativity precession \citep[see Equations (25), (27), and (28) in][]{MHL10}, and $N_<$ is the total number of stars within the radius equal to the semimajor axis of the captured star. Adopting a simple stellar cusp model, i.e., $\rho_*\propto r^{-\alpha}$, we have $N_<=N_{\rm h}(r/r_{\rm h})^{3-\alpha}$, where $r_{\rm h}$ represents the radius within which the mass of stars equals the MBH mass and $N_{\rm h} =M_{\bullet}/m_*=4\times 10^5$ is the total number of field stars within $r_{\rm h}$. Similar to \citet{MHL10}, we also assume a Bahcall--Wolf cusp ($\alpha=7/4$, \citealt{BW76}) unless otherwise stated and correspondingly $r_{\rm h}=2.3{\rm ~pc}$. The variable superscript `(1)' in the above equations means that the time step is one orbital period of the star being investigated. In the time step of $N$ period of the star ($N=\delta t/\delta t_{\rm P}$), the model parameters ($\phi_N$, $\theta_N$, $\sigma_N$) can be obtained from parameters of one period ($\phi_1$, $\theta_1$, $\sigma_1$). For further details of the ARMA model, see \citet{MHL10}. The ARMA model may not capture the exact dynamical physics of the system and give the exact kinematics of each individual star; but for the purpose of our work and the addressing problems, it should be plausible and efficient to be applied here to obtain the evolution of the system in a statistical way. The two-body NR is also taken into account in a way similar to that in \citet{MHL10}. In a time step $\delta t$, the energy change of a captured star due to the NR is given by \begin{equation} \Delta E=\xi E \left(\frac{\delta t}{t_{\rm NR}}\right)^{1/2}, \end{equation} where $\xi$ is an independent normal random variable with zero mean and unit variance, and $t_{\rm NR}$ is the NR timescale in the GC given by \begin{equation} t_{\rm NR}=A_{\rm NR}\left(\frac{M_{\bullet}}{m_*}\right)^2\frac{1}{N_<} \frac{1}{\ln\Lambda}\delta t_{\rm P}. \end{equation} The travel time of those HVSs discovered in the Galactic halo is $\sim 40$-$250$~Myr if HVSs were ejected from the GC \citep{Brown09a}. And recent observations have also shown that the age of the detected HVSs is on the order of $\sim 100$~Myr, which is consistent with the GC origin \citep{Brown12b}. In the model of this paper, we are unifying the formation of both the HVSs and the GC S-stars by the tidal breakup of young stellar binaries originated from the young stellar disk(s) in the GC. In order to be compatible with the above observations, we assume a constant injection rate of stellar binaries over the past $250$~Myr\footnote{Although the majority of the currently detected Wolf--Rayet and O/B type supergiants and giants in the disk are young (with age of $6$~Myr or so; \citealt{Paumard06}), the observations have also shown that there are many B-dwarf stars in the disk region ($0.04-0.5$~pc). For example, \citet{Bartko10} find $59$ B-dwarfs in the disk region and the ages of these dwarfs could be substantially larger than $6$~Myr. Note also that the observational bias on the detection of young but faint stars in the inner parsec is largely uncertain. Thus, current observations do not exclude the existence of less massive stars with ages much longer than those of the detected Wolf--Rayet and O/B type disk stars in the GC.}. We adopt the numerical results of the three-body experiments for each injection model in Section~\ref{sec:num_simu} and calculate the energy and angular momentum evolution for each captured star by using the ARMA model. In these calculations, we also take account of the effects of the limited lifetime of the captured stars on the main sequence and the tidal disruption of those captured stars moving too close to the MBH. We remove those captured stars if they move away from the main sequence or approach the MBH within a distance of $r_{\rm p}<r^{\rm td}$, where $r^{\rm td}=(2M_{\bullet}/m)^{1/3}R_{*}$ and $R_*$ is the stellar radius. Finally, we obtain the present-day semimajor axis and eccentricity distributions of the captured stars, which can be used to be compared to the observational distributions and constrain the models. \begin{figure*} \centering \includegraphics[scale=0.5]{f4.eps} \caption{ Cumulative distributions of the semimajor axis (left panel) and eccentricity (right panel) of the captured stars surviving to the present time. The histograms represent the distributions of the observed GC S-stars \citep{Gillessenetal09}. The solid (red), dotted (blue), dashed (magenta), and dot-dashed (cyan) curves show the results obtained from the Unbd-MS0 model, the Disk-MS0 model, the Disk-TH0 model, and the Disk-TH2 model, respectively. The thick and thin curves represent the distributions of those captured stars with mass $7$-$15M_{\odot}$ and $ 3$-$4M_{\odot}$, respectively. In the right panel, the dot-dot-dashed line represents the cumulative eccentricity distribution proportional to $e_{\rm cap}^{3.6}$ as suggested by the observations \citep{Ghezetal08,Gillessenetal09}. } \label{fig:f4} \end{figure*} The left panel of Figure~\ref{fig:f4} shows the cumulative distributions of $a_{\rm cap}$ of the captured stars surviving to the present time and that of the observed GC S-stars. The thick lines represent the captured stars with mass $\sim 7$-$15M_{\odot}$ surviving to the present time (the simulated GC S-stars), roughly corresponding to the observed GC S-stars, and the thin lines represent the captured stars with mass $\sim 3$-$4M_{\odot}$, roughly corresponding to the captured companions of those HVSs detected in the Galactic halo. Although the lifetime of less massive stars on the main sequence is substantially longer and thus the dynamical evolution time is longer than that of the massive ones, the cumulative distribution of $a_{\rm cap}$ of the light captured stars is only slightly different from that of the massive ones (see the left panel of Figure \ref{fig:f4}). The slope of the $a_{\rm cap}$ distribution is affected most by the $f_{r_{\rm p}}(r_{\rm p,ini})$ distribution. Relatively fewer captured stars are produced in the inner region by the Disk-TH2 model compared with that obtained by the other models, and the fraction of captured stars with semimajor axes $<a_{\rm cap}$ is proportional to $\sima_{\rm cap}^{2.0}$ for the Disk-TH2 model but to $\sim a_{\rm cap}^{1.0-1.5}$ for the other three models. For those models with $\beta=0$, the $a_{\rm cap}$ distributions obtained from the numerical simulations is roughly consistent with the simple estimations from Equation (\ref{eq:fac}) (e.g., the slope is $1$ for $\alpha=-1$ and $\beta=0$ according to Equation (\ref{eq:fac})). For the Disk-TH2 model, however, the $a_{\rm cap}$ distribution seems flatter than the simple expectation, i.e., a slope of $3$ (for $\alpha=-1$ and $\beta=2$). The flatter slope of $f(a_{\rm cap})$ resulted from the Disk-TH2 model may be due to the effect of various mass ratios of the injecting binaries, which is included in the numerical simulations but ignored in deriving Equation (\ref{eq:fac}) (see panel (d) in Figure (\ref{fig:f1})). The Kolmogorov--Simirnov (K-S) tests find the likelihoods of $0.02$, $0.2$, $0.05$, and $0.15$ that the $a_{\rm cap}$ distribution of the observed GC S-stars is the same as that of the simulated GC S-stars for the four models, respectively. As shown in Figure~\ref{fig:f4}, most of the discrepancy between the observational distribution and the simulated one is apparently near the edges of those distributions. Since the Anderson--Darling (A-D) test may be more effective than the K-S test and more sensitive to the distribution edges \citep[see][]{FB12}, we also adopt the A-D test here and find the likelihoods are $0.05$, $0.3$, $0.03$, and $0.32$ for the four models, respectively. These statistical tests suggest that the Disk-MS0 model and the Disk-TH2 model may be more compatible with the observational $a_{\rm cap}$ distribution. For the majority of the simulated GC S-stars ($\sim 7$-$15M_{\odot}$), the relative changes in their energy due to dynamical relaxation after their capture are less than $\sim$30\% and their semimajor axes do not deviate much from the initial values right after their capture. Therefore, the distribution of the semimajor axis of currently observed GC S-stars can provide some information on their ejected companions (see Equations (\ref{eq:dk_b}) and (\ref{eq:acap0})). For those captured stars with mass $\sim 3$-$4M_{\odot}$, however, their relative energy changes can be as large as $1$ mainly because of their longer lifetime and thus longer dynamical evolution time, and Equation (\ref{eq:acap0}) is no longer reliable to provide estimations on the velocity of the ejected companions of those less massive captured stars by using their present-day semimajor axes. The right panel of Figure~\ref{fig:f4} shows the cumulative eccentricity distributions obtained from different models. As seen from Figure~\ref{fig:f4}, the $e_{\rm cap}$ distributions are only slightly different for different injection models because the initial $e_{\rm cap}$ are all close to $1$ in all the models. For those captured stars with mass $\sim 3$-$4M_{\odot}$, their present eccentricities are relatively lower than that of the observed GC S-stars because of their longer dynamical evolution time. For those simulated GC S-stars (with mass $\sim 7$-$15M_{\odot}$) resulted from any of the four injection models, their $e_{\rm cap}$ distribution is similar to that of the observed GC S-stars. The K-S tests find a likelihood of $\sim 0.1-0.6$ that the eccentricity distribution of the observed S-stars is the same as that of the simulated GC S-stars for all the four models. If alternatively adopting the A-D test, then the likelihoods are $\sim 0.004, 0.02, 0.005$, and $0.13$ for the Unbd-MS0 model, the Disk-MS0 model, the Disk-TH0 model, and the Disk-TH2 model, respectively. According to these calculations, the Disk-TH2 model may be more compatible with the observational $e_{\rm cap}$ distribution. We note here that the simulations slightly over-produce the stars with high eccentricities ($e_{\rm cap}$ close to $1$) with respect to the observations because the RR for those captured stars with extremely high eccentricities is quenched by the strong relativistic precession \citep{MHL10}. For this inconsistency, part of the reason might be the observational bias in detecting the GC S-stars, i.e., the stars with high eccentricities are less likely to be detected at $a_{\rm cap}\sim 0.01{\rm ~pc}$ \citep{Schodel03, Weinberg05, MHL10}; and part of the reason might be the limitation of the ARMA model. But the main reason of the inconsistency does not appear to be due to ignoration of the ``bouncing effect'' demonstrated in Figure 7 of \citet{Merritt11}, where the star starting from a low-eccentricity orbit and evolving close to a critical high-eccentricity orbit is then bounced back onto a low-eccentricity orbit due to the suppression of the RR by the fast GR precession at high-eccentricity orbits, as (1) the stars in our model are captured from tidal breakup of binary stars and they initially have eccentricities even higher than the critical eccentricities; (2) during the simulation period, some of the stars have evolved onto low-eccentricity orbits as illustrated by the distribution at the low-eccentricity end in Figure~\ref{fig:f4}, and the simulated stars at the high-eccentricity end are those that evolve relatively slowly; and (3) the suppression of the RR due to the fast GR precession has been modeled in our work, as mentioned before (e.g., see the definition of $t_{\rm prec}$ above). The timescale of the RR process also depends on the mass of the field stars \citep{MHL10,RT96}. To check this dependence, we perform additional simulations by setting the mass of field stars to $m_*=5M_{\odot}$ or $m_*=20M_{\odot}$ but with the total mass of the field stars fixed. According to the results of these simulations, we find that the simulated GC S-stars are on orbits with too high eccentricities compared with the observational ones if $m_*=5M_{\odot}$, or on orbits with too low eccentricities if $m_*=20M_{\odot}$. In the above calculations, a Bahcall--Wolf cusp for the background stellar system in the GC is adopted. However, recent observations suggested that the background stellar distribution may be core-like rather than cusp-like (e.g., \citealt{Doetal09}). Similarly as done in \citet{MHL10}, we also adopt $\alpha=0.5$ to mimic the effect of a core-like distribution, and the perturbation on the GC ``S-stars'' is assumed to be dominated by main-sequence stars (e.g., see \citealt{Antonini12}). In such a model, we find that the resulted S-stars on highly eccentric orbits with smaller pericenter distances are relatively more than those obtained from the cuspy model. The reason is that the RR is less efficient in the core-like stellar distribution, and thus the evolution of the eccentricities of GC ``S-stars'' is slower. The timescale of the RR process becomes much longer at the distance of the disks. Stars in this region are less affected by the relaxation processes and may well preserve some of their initial orbital configurations. Observations find many B-dwarfs in disk regions, with high eccentricities and more extended spatial distribution than disk stars. The resulted eccentricity--distance distribution of these stars by the RR compared with the observations may provide useful constraints on the formation of the GC S-stars \citep{Perets10}. Our simulations also produce many B-dwarfs in the disk region and their radial distribution is similar to the initial input ones for the injecting binaries. However, there should also exist B-dwarfs in the disk region that are initially formed as single stars and the fraction of these stars is not clear yet. A detailed dynamical study for the B-dwarfs in the disk region is complicated and beyond the scope of this paper. \section{Ejected HVSs in the Galactic bulge and halo}\label{sec:HVSs} The ejected stars move away from the GC after the breakup of their progenitor binaries and their velocities are gradually decelerated in the Galactic gravitational potential. Some of them are unbound to the Galactic potential and can travel to the Galactic halo and may appear as the detected HVSs if their main-sequence lifetime is long enough compared to the travel time; while others with lower ejecting velocity may return to the GC. To follow the subsequent motion of the ejected components, we adopt the Milky Way potential model given by \citet{Xue08}, which involves four components, including the contributions from the central MBH, the Galactic bulge, the Galactic disk, and the Galactic halo, i.e., \begin{equation} \Phi=\Phi_{\rm BH}+\Phi_{\rm bulge}+\Phi_{\rm disk}+ \Phi_{\rm halo}, \end{equation} where \begin{equation} \Phi_{\rm BH} =-GM_{\bullet}/r, \end{equation} \begin{equation} \Phi_{\rm bulge}=-\frac{GM_{\rm bulge}}{r+r_{\rm bulge}}, \end{equation} \begin{equation} \Phi_{\rm disk} =-\frac{GM_{\rm disk}(1-e^{-r/b})}{r}, \end{equation} \begin{equation} \Phi_{\rm halo} =-\frac{4\pi G\rho_{\rm s} r^3_{\rm vir}}{c^3r} \ln(1+\frac{cr}{r_{\rm vir}}), \end{equation} respectively. The model parameters for the last three components are $M_{\rm bulge}=1.5\times10^{10}M_{\odot}$, $M_{\rm disk}=5\times10^{10}M_{\odot}$, the core radius $r_{\rm bulge}=0.6\rm ~kpc$, the scale length $b=4\rm ~kpc$, and $\rho_{\rm s}=\frac{1}{3}\frac{c^3\rho_{\rm c}\Omega_{\rm m}\Delta_{\rm vir}}{\ln(1+c)-c/(1+c)}$, where $\rho_{\rm c}$ is the cosmic critical density, $\Delta_{\rm vir}=200$, $\Omega_{\rm m}$ is the cosmic fraction of matter, the virial radius $r_{\rm vir}=267\rm ~kpc$, and the concentration $c=12$. The bulge, the disk, and the halo potentials adopted here are all spherical. If adopting non-spherical potentials, i.e., a triaxial bulge/halo and a flattened disk potential, the bending effect due to the non-spherical component on the trajectories of ejected stars is important only for those with $v_{\infty} \mathrel{\mathpalette\oversim<} 400{\rm km\,s^{-1}}$ on a timescale of $\mathrel{\mathpalette\oversim>} 500$~Myr, but it is negligible for HVSs with relatively high speeds \citep[see][]{YM07}. Note that the radial distribution of the ejected stars surviving to the present time (and correspondingly the predicted number of the detectable HVSs) may be slightly different if adopting a different Galactic potential model. The total number of detectable HVSs depends directly not only on how many stellar binaries can be injected into the immediate vicinity of the MBH, but also on the lifetime of these stars and the detailed settings on the IMF, semimajor axis, and periapsis of the injecting stellar binaries (see Section~\ref{sec:num_simu}). If the stellar binaries are injected from a far away region, the injection rate can be estimated through the loss-cone theory \citep{YT03,Perets09c}; if the injecting stellar binaries originated from central stellar disks \citep{Luetal10,Zhang10}, the injection rate is difficult to estimate as the mechanism responsible for it is still not clearly understood (cf.\ \citealt{MLH09}). In principle, it is plausible to observationally calibrate the injection rate by the numbers of the detected HVSs and GC S-stars. But this calibration becomes complicated if considering of the uncertainties in the settings of the distributions $f_{a_{\rm b}}(a_{\rm b,ini})$, $f_{r_{\rm p}}(r_{\rm p,ini})$, and IMF of the injecting binaries, etc. \begin{deluxetable*}{lccccccccc} \tablecaption{The Number Ratio of the Simulated $3$-$4M_{\odot}$ HVSs to the Simulated GC S-stars} \tablehead{ \colhead{Model} & \colhead{$\gamma$} & \colhead{$\beta$} & \colhead{$\frac{N_{{\rm HVS}}^{{\rm tot}}}{N_{{\rm cap}}^{{\rm tot}}}$} & \colhead{$\frac{F^{{\rm lt}}_{{\rm HVS}}}{F^{{\rm lt}}_{{\rm cap}}}$} & \colhead{$F^{\rm td}_{{\rm cap}}$} & \colhead{$F_{\rm HVS}^{\rm obs,rf}$} & \colhead{$F_{{\rm cap}}^{\rm obs}$} & \colhead{$\frac{N_{\rm HVS}^{\rm obs,rf}}{N_{{\rm cap}}^{\rm obs}}$}} \startdata Unbd-MS0 & -2.7 & 0 & 3.0 & 8.8 & 0.59 & 0.27 & 0.51 & 27 \\ Disk-MS0 & -2.7 & 0 & 1.8 & 8.9 & 0.51 & 0.15 & 0.41 & 12 \\ Disk-TH0 & -0.45 & 0 & 0.22 & 10 & 0.47 & 0.19 & 0.54 & 1.4 \\ Disk-TH2 & -0.45 & 2 & 0.27 & 9.9 & 0.34 & 0.06 & 0.21 & 1.3 \\ Disk-IM0 & -1.6 & 0 & 0.61 & 9.5 & 0.50 & 0.16 & 0.47 & 4.0 \\ Disk-IM2 & -1.6 & 2 & 0.75 & 9.4 & 0.22 & 0.01 & 0.03 & 2.8 \enddata \label{tab:t3} \tablecomments{ The $N_{\rm HVS}^{{\rm tot}}$ and $N_{{\rm cap}}^{{\rm tot}}$ represent the total number of the ejected stars with mass $3$-$4M_{\odot}$ and the total number of the captured stars with mass $7$-$15M_{\odot}$ that are generated by the tidal breakup of stellar binaries in the GC for each model, respectively; the $F^{{\rm lt}}_{\rm HVS}$ and $F^{{\rm lt}}_{{\rm cap}}$ denote the fraction of the ejected and captured stars that still remain on the main sequence of their stellar evolution at the end of our simulations; the $F^{\rm td}_{{\rm cap}}$ denotes the fraction of the captured stars that have been tidally disrupted by the central MBH before the end of our simulations; the $F_{\rm HVS}^{\rm obs,rf}$ denotes the fraction of the ejected stars appear as the detected HVSs, where an ejected star is taken as a {\it detectable} HVS if its heliocentric radial velocity in the Galactic rest frame is $|v_{\rm rf}| > 275{\rm km\,s^{-1}}$, its velocity at infinity is $v_{\infty} >750{\rm km\,s^{-1}}$, and its distance from the GC is in the range of $40$-$130\rm ~kpc$; $F_{{\rm cap}}^{\rm obs}$ denotes the fraction of the captured stars that are within radii $\mathrel{\mathpalette\oversim<} 4000{\rm AU}$ from the MBH; and the $N_{\rm HVS}^{\rm obs,rf}$ is the total number of those {\it detectable} HVSs with mass $3$-$4M_{\odot}$ in the Galactic halo and $N_{{\rm cap}}^{\rm obs}$ is the simulated number of the captured stars with mass $7$-$15M_{\odot}$ in the GC, which correspond to the observed ones. } \end{deluxetable*} \begin{deluxetable*}{lccccccccccccccc} \tablewidth{18.5cm} \tabletypesize{\tiny} \tablecaption{Predicted Numbers in Different Models} \tablecolumns{2} \tablehead{ \multirow{2}{*}{Model} & \multirow{2}{*}{$\gamma$} & \multirow{2}{*}{$\beta$} & \colhead{injection rate} & \colhead{} & \multicolumn{3}{c}{$3$-$4M_{\odot}$} & \colhead{} & \multicolumn{3}{c}{$4$-$7M_{\odot}$}& \colhead{} & \multicolumn{3}{c}{$7$-$15M_{\odot}$} \\ \cline{6-8} \cline{10-12} \cline{14-16} & & & $(10^{-5}\,{\rm yr^{-1}})$& & $N_{\rm HVS}^{\rm obs,rf}$ & $N_{\rm HVS}^{\rm obs,pm}$ & $N_{{\rm cap}}^{\rm obs}$ & & $N_{\rm HVS}^{\rm obs,rf}$ & $N_{\rm HVS}^{\rm obs,pm}$ & $N_{{\rm cap}}^{\rm obs}$ & & $N_{\rm HVS}^{\rm obs,rf}$ & $N_{\rm HVS}^{\rm obs,pm}$ & $N_{{\rm cap}}^{\rm obs}$ } \startdata Unbd-MS0 &-2.7 & 0 & 13 (2.2) & &504 (79)&189 (30)& 86 (13)& &178 (28)& 58 (9) &33 (5) & & 60 (9) & 33 (5) & 17 (3)\\ Disk-MS0 &-2.7 & 0 & 7.0 (2.7) & &177 (79)& 66 (29)& 82 (36)& & 62 (28)& 21 (9) & 32 (14)& & 29 (13)& 17 (8) & 17 (8)\\ Disk-TH0 &-0.45 & 0 & 1.8 (5.8) & &23 (79)&8 (27)& 11 (37)& & 28 (97)& 8 (30) & 11 (39)& &29 (103)& 19 (66)& 17 (60)\\ Disk-TH2 &-0.45 & 2 & 6.2 (23) & &18 (79)&7 (31)& 8 (34)& & 16 (71)& 5 (23) & 9 (40)& & 20 (87)& 15 (64)& 17 (75)\\ Disk-IM0 &-1.6 & 0 & 3.1 (3.6) & &61 (79)& 22 (29)& 28 (36)& & 37 (48)& 12 (15) & 18 (23)& & 29 (38)& 18 (23)& 17 (22)\\ Disk-IM2 &-1.6 & 2 & 60 (100) & &48 (79)& 21 (35)& 32 (52)& & 24 (39)& 7 (11) & 18 (29)& & 20 (33)& 14 (23)& 17 (28) \enddata \label{tab:t4} \tabletypesize{\footnotesize} \tablecomments{ The numbers of the simulated {\it detectable} HVSs and the captured stars surviving in the GC at the present time in different mass ranges, obtained from different models. The injection rate of stellar binaries is assumed to be a constant over the past $250$~Myr, which enables the production of $17$ simulated GC S-stars (or $79$ unbound $3\sim4M_{\odot}$ HVSs (numbers in the brackets)). The $N_{\rm HVS}^{\rm obs,rf}$ denotes the number of HVSs with $|v_{\rm rf}|>275{\rm km\,s^{-1}}$ and $v_{\infty}>750{\rm km\,s^{-1}}$; and the $N_{\rm HVS}^{\rm obs,pm}$ denotes the number of the HVSs with proper motion $\ge 5{\rm mas\,yr^{-1}}$ in the heliocentric rest frame and $v_{\infty}>750{\rm km\,s^{-1}}$. In order to compare to the observations, a simulated HVS with mass of $3$-$4M_{\odot}$ is counted as a {\it detectable} HVS additionally if its distance to the GC is in the range of $40-130\rm ~kpc$. The $N_{{\rm cap}}^{\rm obs}$ is the total number of the captured stars surviving in the GC with present-day $a_{\rm cap}\mathrel{\mathpalette\oversim<} 4000{\rm AU}$ in the simulations. } \end{deluxetable*} \subsection{The Numbers of the HVSs/GC S-stars and Their Number Ratio}\label{subsec:number} The $3$-$4M_{\odot}$ HVSs detected in the Galactic halo and the GC S-stars should be linked to each other under the working hypothesis of this paper. The total numbers of the simulated $3$-$4M_{\odot}$ HVSs and $7$-$15M_{\odot}$ GC S-stars depend not only on the injection rate of stellar binaries but also on the detailed settings on the injection models. However, the number ratio of the simulated $3$-$4M_{\odot}$ HVSs to the simulated $7$-$15M_{\odot}$ GC S-stars may depend only on the details of the injection models described in Section~\ref{sec:num_simu}, but not on the injection rate of stellar binaries. Any viable model should produce a number ratio compatible with the observations on the HVSs and the GC S-stars, and thus this number ratio may provide important constraints on the models. We obtain both the numbers of the simulated HVSs and the captured stars surviving to the present time and their number ratio by Monte Carlo simulations as follows. First, we obtain the total number of initially captured (or ejected) stars, $N_{{\rm cap}}^{{\rm tot}}$ (or $N_{{\rm HVS}}^{{\rm tot}}$), with mass in the ranges of $3$-$4M_{\odot}$, $4$-$7M_{\odot}$, and $7$-$15M_{\odot}$, respectively. To do this, we assume a constant injection rate of binaries and randomly set the injection events over the past $250$~Myr, and for each injection event we randomly assign it to a three-body experiment conducted in Section~\ref{sec:num_simu} and adopt the results from the experiment. Second, we consider the limited lifetime of each ejected and captured star, the motion of each ejected star in the Galactic potential, and the dynamical evolution of each captured star, and then obtain the fraction of the ejected stars ($F_{\rm HVS}^{{\rm lt}}$) that still remain on the main sequence at the present time or the similar fraction for the captured stars ($F_{{\rm cap}}^{{\rm lt}}$), and the fraction of the captured stars ($F_{{\rm cap}}^{\rm td}$) that have already been tidally disrupted until the present time. Third, we consider the kinematic selection criteria and obtain the fractions of those ejected and captured stars according to given observational selection criteria, i.e., $F_{\rm HVS}^{\rm obs}$ and $F_{{\rm cap}}^{\rm obs}$, respectively. The selection criteria are similar to those adopted in selecting the detected HVSs and GC S-stars, i.e., an ejected star is labeled as a {\it detectable} HVS if its heliocentric radial velocity in the Galactic rest frame is $|v_{\rm rf}| >275{\rm km\,s^{-1}}$ or its proper motion in the heliocentric rest frame is $\ge 5{\rm mas\,yr^{-1}}$, and its velocity at infinity is $v_{\infty} >750{\rm km\,s^{-1}}$, and captured stars with semimajor axis $\mathrel{\mathpalette\oversim<} 4000{\rm AU}$ are counted as {\it detectable} GC S-stars. For those simulated HVSs with mass $ 3$-$4M_{\odot}$, we put an additional cut on their distances from the GC, i.e., from $40$ to $130\rm ~kpc$, in order to compare them to current observations. Finally, we obtain the number ratio of the {\it detectable} HVSs to the {\it detectable} GC S-stars for each model as \begin{equation} \frac{N_{\rm HVS}^{\rm obs}}{N_{{\rm cap}}^{\rm obs}}= \frac{N_{{\rm HVS}}^{{\rm tot}}}{N_{{\rm cap}}^{{\rm tot}}}\times \frac{F^{\rm lt}_{{\rm HVS}}}{F^{{\rm lt}}_{{\rm cap}}}\times \frac{F_{\rm HVS}^{\rm obs}}{(1-F^{\rm td}_{{\rm cap}}){F_{{\rm cap}}^{\rm obs}}}. \end{equation} The simulation results are listed in Tables~\ref{tab:t3} and~\ref{tab:t4} for each model. Here we comment on a few factors that affect the predicted numbers of the simulated {\it detectable} HVSs and GC S-stars and consequently the number ratio of these two populations. (1) The difference between the lifetime of the simulated HVSs and GC S-stars: the detected HVSs are in the mass range of $\sim 3$-$4M_{\odot}$, which are substantially smaller than that of the observed GC S-stars ($\sim 7$-$15M_{\odot}$). The lifetime difference leads to an enhancement in the number ratio of the {\it detectable} HVSs to the GC S-stars roughly by a factor of $10$ under the assumption of a constant rate of injecting stellar binaries into the vicinity of the MBH over the past $250$~Myr. (2) The place where the stellar binaries are originated: in the Unbd-MS0 model, relatively more stars with high velocities (e.g., $>1000{\rm km\,s^{-1}}$) are generated than those in the other models. (3) The distribution of pericenter distance of those injecting stellar binaries, which is related to the speed of the migration or diffusion process of those binaries to the immediate vicinity of the central MBH: a change in this distribution may result in either a significant increase or decrease in both the number of HVSs and that of captured stars, but their number ratio is not affected much. The complete survey conducted by \citet{Brown09a} has detected $14$ unbound HVSs with mass $\sim 3$-$4M_{\odot}$ in a sky area of $\sim 7300$~deg$^2$,\footnote{ One sdO type star with mass $\sim 1M_{\odot}$ and another massive HVS in the southern hemisphere with mass $\sim 9M_{\odot}$ in the survey are not included in the number.} and the total number of similar HVSs in the whole sky should be $\sim 79\pm21$.\footnote{Considering of the new results on searching HVSs reported by \citet{Brown12}, this number could be slightly higher, i.e., $95\pm 23$. And if assuming that the detected HVSs were originated from two disk-like stellar structures, i.e., the CWS disk plane and the Narm plane, as suggested by \citet{Luetal10}, the expected total number of HVSs with mass $\sim 3$-$4M_{\odot}$ is $75\pm28$. The number ratios of the detected HVSs to the GC S-stars become $\sim 2.2$-$5.0$.} Observations have revealed $17$ GC S-stars within a distance of $\sim 4000{\rm AU}$ from the central MBH. The number ratio of the detected HVSs to the GC S-stars is $\sim3.4$-$5.9$. The number ratio resulted from any of the top four injection models listed in Table~\ref{tab:t2} is inconsistent with the observational ones. Both the Unbd-MS0 model and the Disk-MS0 model give a number ratio substantially larger than that inferred from observations, while the Disk-TH0 model and the Disk-TH2 model give too small number ratios. \citet{Figer99} suggest that the IMF of young star clusters in the GC, i.e., the Arches cluster, may be top-heavy and the IMF slope is $\gamma\sim -1.6$, although some later studies argued that the IMF of Arches cluster may be still consistent with the Salpeter one by assuming continuous star formation \citep[e.g.,][]{Lockmann10}. Considering of this, we adopt an IMF with a slope of $\gamma=-1.6$ and perform two additional injecting models, i.e., ``Disk-IM0'' and ``Disk-IM2'', as listed in Tables~\ref{tab:t3} and \ref{tab:t4}. Our calculations show that the number ratios produced by the two models are close to the observational ones. Obviously, the number ratio of the simulated HVSs to GC S-stars is significantly affected by the adopted IMF. Adopting a steeper IMF may lead to a larger number ratio. The velocity distribution of the detected HVSs suggests a slow migration/diffusion of stellar binaries into the immediate vicinity of the central MBH \citep[see detailed discussions in][]{Zhang10}. The Disk-IM2 model can produce a velocity distribution similar to the observational ones, and the large $\beta$ adopted in this model also suggests a slow migration/diffusion of binaries into the vicinity of the central MBH. However, all the other models appear to generate too many HVSs at the high-velocity end and thus a too flat velocity distribution compared with the observational ones. This inconsistency could be due to many factors. For example, (1) the velocity distribution of the detected HVSs could be biased due to either the small number statistics and/or the uncertainties in estimating the observational selection effects; (2) the uncertainties in the initial settings of the injecting binaries could also lead to some change in the velocity distribution, e.g., the velocity distribution may be steeper if the semimajor axis distribution of stellar binaries is log-normal, rather than follow the \"{O}pik law \citep{Sesana07,Zhang10}; and (3) the injection of binaries to the vicinity of the central MBH may be quite different from the simple models adopted in this paper because of the complicated environment of the very central region, e.g., the possible existence of a number of stellar-mass BHs or an intermediate-mass black holes (BHs) within $100{\rm AU}$. For each model, the injection rate of stellar binaries can be calibrated to produce the numbers of the observed HVSs and GC S-stars. Under the initial settings for each model described in Section~\ref{sec:num_simu}, the numbers of the simulated HVSs and the GC S-stars, similar to the detected ones, are listed in Table~\ref{tab:t4}. The calibrated injection rates are also listed in Table~\ref{tab:t4} for each model and they are roughly on the order of $\sim10^{-3}$ to $10^{-5}~{\rm yr}^{-1}$. However, one should be cautious about that this injection rate depends on the initial settings on the distributions of the semimajor axes and the periapses of the injecting binaries. According to our simulations, many of the injected binaries are not tidally broken up, and the breakup rate is a factor of $3$-$6$ times smaller than the injection rate for those models studied in this paper, i.e., about a few times $10^{-4}$ to $10^{-5}~{\rm yr}^{-1}$, which is consistent with the estimates by \citet{Bromley12}. \subsection{The Ejected Companions of the GC S-stars} \begin{figure*} \centering \includegraphics[scale=0.5]{f5.eps} \caption{ Cumulative distributions of the Galactocentric distance $R_{\rm GC}$ (panel (a)), the velocity in the Galactocentric rest frame (panel (b)), the heliocentric radial velocity in the Galactic rest frame (panel (c)), and the proper motion in the heliocentric rest frame (panel (d)), of those simulated unbound HVSs ($v_\infty>750{\rm km\,s^{-1}}$) that were initially associated with the ``GC S-stars''. The solid (red), dotted (blue), dashed (magenta), and dot-dashed (cyan) curves are for the Unbd-MS0 model, the Disk-MS0 model, the Disk-TH0 model, and the Disk-TH2 model, respectively. } \label{fig:f5} \end{figure*} \begin{figure} \centering \includegraphics[scale=0.5]{f6.eps} \caption{ Spatial distribution of the simulated unbound HVSs with mass $7$-$15M_{\odot}$ and $v_\infty>750{\rm km\,s^{-1}}$ for the Disk-IM2 model. The distribution is expressed by a Hammer-Aitoff projection in the Galactic coordinates. In the top panel, the positions of the HVSs are projected to infinity from the GC. The solid red and blue curves represent the CWS and NARM disk planes, respectively, which are also projected to infinity. The region above the green curve shows the area surveyed by \citet{Brown09a} in the northern hemisphere. The open red and blue circles represent the unbound HVSs originated from binary stars on the CWS and NARM disk planes, respectively. For illustration purpose, all the $34$ simulated HVSs are shown in the top panel (see Table \ref{tab:t4}). The injection rates are assumed to be the same for the injections from the CWS and the NARM planes and thus the number of HVSs associated with the two planes are also the same. In the bottom panel, the positions of the HVSs are not projected to infinity. The solid red curves and the solid green curve are the same as those for the top panel. The red crosses and blue plus symbols represent those unbound HVSs with $|v_{\rm rf}|\ge 275{\rm km\,s^{-1}}$ injected from the CWS and NARM disk planes, respectively. The red open diamonds and the blue open squares represent the unbound HVSs with proper motion $\mu>5{\rm mas\,yr^{-1}}$ injected from the CWS and the NARM disk planes, respectively. } \label{fig:f6} \end{figure} The HVSs discovered in the Galactic halo are typically in the mass range $\sim 3$-$4M_{\odot}$. HVSs with other masses should also be populated in the Galactic bulge and halo. In this Section, we investigate the properties of the simulated HVSs with mass $\sim 7$-$15M_{\odot}$, which are most likely to be the companions of the ``GC S-stars''. Panel (a) of Figure~\ref{fig:f5} shows the cumulative distribution of the Galactocentric distances of the high-mass ($\sim 7$-$15M_{\odot}$) HVSs. The predicted numbers of these HVSs from different models are listed in Table~\ref{tab:t4}. The total number of the simulated unbound HVSs, as companions of the population of the GC S-stars, is $\sim 20$-$30$ according to the Disk-IM2 model, which is able to reproduce the numbers of the observed HVSs and GC S-stars. The majority of these unbound HVSs are at distances of a few to a few tens $\rm ~kpc$ from the GC, which are much closer to the GC than the detected $3$-$4M_{\odot}$ HVSs mainly because a high-mass star has a shorter main-sequence lifetime and thus the distance it can travel within the lifetime is small. The close distances of these HVSs from the GC suggest that the velocity vectors of many of them are not along our line of sight. Their three-dimensional velocities range from $500{\rm km\,s^{-1}} $ to $2000{\rm km\,s^{-1}}$ in the Galactocentric rest frame (see panel (b) in Figure \ref{fig:f5}); and their heliocentric radial velocities in the Galactic rest frame range from $-500{\rm km\,s^{-1}}$ to $1500{\rm km\,s^{-1}}$ (see panel (c) in Figure \ref{fig:f5}), where the negative and positive velocities represent moving toward and away from the Sun, respectively. Compared with the ejection velocities of the $3$-$4M_{\odot}$ HVSs, those of the $7$-$15M_{\odot}$ HVSs are relatively higher because of their higher mass (see Equation (\ref{eq:vhvs})). Most of the HVSs have proper motions in the heliocentric rest frame as large as ${\rm mas\,yr^{-1}}$ to a few tens ${\rm mas\,yr^{-1}}$ (see panel (d) in Figure~\ref{fig:f5}). These HVSs are bright enough to be detected at a distance less than a few tens $\rm ~kpc$ by future telescopes, such as, the {\it Global Astrometric Interferometer for Astrophysics} spacecraft ({\it Gaia}), and their proper motions are also large enough to be measured. Figure~\ref{fig:f6} shows the sky distribution of the simulated HVSs with mass$\sim 7$-$15M_{\odot}$, which may represent the ejected companions of the ``GC S-stars'', in the Galactic coordinates by a Hammer Aitoff projection. In the top panel of Figure~\ref{fig:f6}, the positions of the HVSs are projected to infinity from the GC, which are consistent with being located close to the CWS disk plane and the Narm plane (also projected to infinity) as expected. In the bottom panel of Figure~\ref{fig:f6}, the positions of the HVSs are not projected to infinity. As seen from the bottom panel, the simulated HVSs with mass $\sim 7$-$15M_{\odot}$ lie in the area below the projected curves of the CWS disk plane and the Narm plane, and most of these high-mass HVSs reside out of the area surveyed by \citet[][]{Brown09b}. Our calculations show that less than $10\%$ of the simulated unbound HVSs with mass $7\sim15M_{\odot}$ are located in the survey area. This may be the reason that none of those high-mass HVSs, possibly the companions of the ``GC S-stars'', has been discovered in the survey area. If the HVSs are initially originated from the CWS disk and the Narm plane, the HVSs with high radial velocities are also relatively rare in the direction close to the disk normals, i.e., $(l,b) = (311{^{\circ}}, -14{^{\circ}})$ and ($176{^{\circ}}, -53{^{\circ}}$) while the HVSs with high proper motions ($\sim 20{\rm mas\,yr^{-1}}$) is relatively numerous in that direction because the velocity vectors of HVSs are close to be perpendicular to the disk normals. Surveys of HVSs in the southern sky with SkyMapper and others may find such massive HVSs, as the possible companions of the ``GC S-stars'', and provide crucial evidence for whether those GC S-stars are produced by the tidal breakup of stellar binaries. Note that the spatial distribution of the HVSs discussed in this section is directly related to the assumption that the injecting stellar binaries are originated from two disk-like stellar structures similar to the CWS disk in the GC. However, the other properties of the HVSs or the GC S-stars discussed in this paper are affected by whether the injecting stellar binaries are bound to the central MBH or not, but not affected by whether they are initially on the CWS disk plane or not. \section{The innermost captured star}\label{sec:innermost} The Unbd-MS0, Disk-MS0, Disk-TH0, Disk-TH2, Disk-IM0, and Disk-IM2 models roughly produce $147$, $111$, $21$, $18$, $52$, and $46$ captured stars surviving to the present time and with mass in the range of $3$-$7M_{\odot}$, less massive than that of the GC S-stars, within a distance of $\sim 4000{\rm AU}$ from the MBH (see Table \ref{tab:t4} and Section~\ref{sec:HVSs}). The above numbers are obtained by calibrating the injection rate of the stellar binaries over the past $250$~Myr to generate $17$ simulated GC S-stars similar to the observational number. The captured stars with mass in the range of $ 3$-$7M_{\odot}$ could be detected by the next generation telescopes, e.g., the Thirty Meter Telescope (TMT) or the European Extremely Large Telescope (E-ELT). These low-mass stars are potentially important probes for testing the GR effects near an MBH, if they are closer to the central MBH than S2. In this section, we estimate the probability distribution of the innermost captured low-mass stars ($\sim 3$-$7M_{\odot}$) by Monte Carlo realizations based on the calibrated injection rate. The left panel of Figure~\ref{fig:f7} shows the probability distributions of the semimajor axis of the innermost captured star with mass $\sim 3$-$7M_{\odot}$ resulted from different injection models. For those models adopting $\beta=0$, the resulted innermost captured star is typically on an orbit with semimajor axis $\sim 300 {\rm AU}$, and the probability that its semimajor axis is less than that of S2 is $\sim 99\%$. For the other models adopting $\beta=2$, the resulted innermost captured star is on an orbit with semimajor axis of $\sim 300$-$1500 {\rm AU}$ and the probability that its semimajor axis is smaller than that of S2 is $\sim 60\%$-$70\%$. The probability to capture a star within the orbit of S2 is larger for the $\beta=0$ models than for the $\beta=2$ models. The reason is that relatively more stellar binaries can be injected into the immediate vicinity of the central MBH and thus more stars can be captured onto orbits with smaller semimajor axes (see Table \ref{tab:t4}) in the models adopting $\beta=0$ than that adopting $\beta=2$. We conclude that the probability of a less massive star ($3$-$7M_{\odot}$) existing within the S2 orbit is at least $61\%$ and can be up to $99\%$, which may be revealed by future observations and then offer important tests to general relativity. The right panel of Figure~\ref{fig:f7} shows the probability distribution of the pericenter distance of the innermost captured low-mass star ($\sim 3$-$7M_{\odot}$) resulted from different injection models. For the injection models adopting $\beta=0$, the pericenter distance distribution is concentrated within $50{\rm AU}$; while for the other models adopting $\beta=2$, the expected pericenter distance is broadly distributed over 10--200\,AU. Nevertheless, the probability that the pericenter distance of the innermost captured star with mass $\sim 3$-$7M_{\odot}$ is less than that of S2 (and S14) is still significant, i.e., $\mathrel{\mathpalette\oversim>} 55\%$ (or $38\%$). The innermost captured star may have its semimajor axis and pericenter distance both significantly smaller than those of S2, therefore, the GR effects on its orbit may be much more significant than that on S2. If taking into account the captured stars with even lower masses, e.g., $1M_{\odot}$, the number of the expected captured stars surviving to the present time becomes much larger, especially for those models with large $\gamma$. For example, the numbers of the captured stars with mass 1--7$M_{\odot}$ are $907$, $841$, $49$, $39$, $214$, and $117$ for the six models, respectively. For those lower mass captured stars, the semimajor axis and the pericenter of the innermost one could be even closer to the central MBH. Some stars may be transported to the vicinity of the central MBH by some mechanisms other than the tidal breakup of stellar binaries. It is possible that some of these stars, with their origins different from the captured stars discussed above, exist within the S2 orbit, but which is beyond the scope of the study in this paper. \begin{figure*} \centering \includegraphics[scale=0.5]{f7.eps} \caption{ Probability distributions of the semimajor axis (left panel, $P(a_{{\rm cap}})$) and the pericenter distance (right panel, $P(r_{\rm p,cap})$) of the innermost captured star with mass in the range $\sim 3$-$7M_{\odot}$ (lower than the masses of the GC S-stars) at the present time. The solid (red), dotted (blue), short-dashed (magenta), dot-dashed (cyan), triple -dot-dashed (green), and long-dashed (yellow) lines show the results obtained from the Unbd-MS0, Disk-MS0, Disk-TH0, Disk-TH2, Disk-IM0, and Disk-IM2 models, respectively. Note here that these probability functions rely on the estimation of the numbers of the simulated detectable captured stars (see Section~\ref{sec:HVSs}). The estimates are obtained by calibrating the injection rate of stellar binaries over the past $250$~Myr to generate the same number (17) of simulated GC S-stars surviving to the present time as that of the observed ones. For reference, the position of S2 (or S14) is labeled in the figure. The probabilities that the innermost star is located within the S2 orbit (i.e., $a_{\rm cap}<1000$AU) are $0.99$, $0,99$, $0.99$, $0.61$, $0.99$, and $0.73$ for the six models, respectively. And the probabilities that the pericenter distance of the inner most star is smaller than that of S2 (S14), i.e., 120\,AU (76\,AU), are $0.86$ ($0.83$), $0.81$ ($0.75$), $0.78$ ($0.70$), $0.55$ ($0.38$), $0.80$ ($0.70$), and $0.62$ ($0.46$) for the six models, respectively. } \label{fig:f7} \end{figure*} \section{Conclusions}\label{sec:conclusion} In this paper, we investigate the link between the GC S-stars and the HVSs discovered in the Galactic halo under the hypothesis that they are both the products of the tidal breakup processes of stellar binaries in the vicinity of the central MBH. We perform a large number of the three-body experiments and the Monte Carlo simulations to realize the tidal breakup processes of stellar binaries by assuming a continuous binary injection rate over the past $250$~Myr, and adopting several sets of initial settings on the injection of binaries. After the tidal breakup of a binary, we follow the dynamical evolution of the captured components in the GC by using the ARMA model (see \citealt{MHL10}), which takes into account both the RR and NR processes, and we also trace the kinematic motion of the ejected component in the Galactic gravitational potential. The properties of the ejected and captured components of the tidally broken-up binaries are naturally linked to each other as they are both the products of tidal breakup of binaries. For those HVSs discovered in the Galactic halo with mass $\sim 3$-$4M_{\odot}$ and $v_{\infty} \sim 700$-$1000{\rm km\,s^{-1}}$, their companions are expected to be captured onto orbits with semimajor axis in the range $\sim 1000$-$8000{\rm AU}$; for the observed GC S-stars with semimajor axis $\sim 1000$-4000${\rm AU}$ in the GC, their companions are expected to be ejected out to the Galactic bulge and halo with $v_{\infty} \sim 500$-$2000{\rm km\,s^{-1}}$. The energy of the captured stars evolves with time because of their dynamical interactions with the environment. For the captured stars with mass $\sim 7$-$15M_{\odot}$, the differences between their present-day energy and their initial ones are no more than $30\%$; for the captured stars with mass $\sim 3$-$4M_{\odot}$, however, the difference can be by order of unity. Therefore, the current semimajor axis distribution of the GC S-stars may provide a good estimation on the velocity distribution of their ejected companions (e.g., Equation (\ref{eq:acap0})), but that of the captured stars with mass $\sim 3$-$4M_{\odot}$ does not. The eccentricities of the ``GC S-stars'' ($\sim 7$-$15M_{\odot}$) are close to $1$ right after the capture and may evolve to low values, and the eccentricity distribution of these simulated GC S-stars at the present time could be statistically compatible with the observational ones of the GC S-stars attributed to the RR processes. For those captured stars with mass $\sim 3$-$4 M_{\odot}$, their eccentricities can evolve to even lower values at the present time compared with the high-mass GC S-stars ($\sim 7$-$15M_{\odot}$) because they interact with the environment for a longer time. To reproduce both the numbers of the detected HVSs and GC S-stars, the injection rate of binaries need to be on the order of $10^{-4}$ to $10^{-5}{\rm yr}^{-1}$ and the IMF of the primary components is required to be somewhat top-heavy with a slope of $\sim 1.6$. For the injection models that can reproduce the observational results on both the GC S-stars and the HVSs, including the distributions of the semimajor axes and eccentricities of the GC S-stars, the spatial and velocity distributions of the detected HVSs, and the number ratio of the HVSs to the GC S-stars, the expected number of the $\sim 3$-$7M_{\odot}$ captured companions is $\sim 50$ within a distance of $\sim 4000{\rm AU}$ from the central MBH. Future observations on the low-mass captured stars may provide a crucial check on whether the GC S-stars are originated from the tidal breakup of stellar binaries. The companions of the HVSs, which are captured by the central MBH, are usually less massive than that of the GC S-stars ($\sim 7$-$15M_{\odot}$). The semimajor axis of the innermost captured star with mass $\sim 3$-$7M_{\odot}$ is $\sim 300$-$1500{\rm AU}$, and the probability that it is smaller than that of S2 is $\sim 70\%$-$90\%$ for the $\beta=2$ models and $\sim 99\%$ for the $\beta=0$ models. The pericenter distance of the innermost captured star with mass $\sim 3$-$7M_{\odot}$ is $\sim 10$-$200{\rm AU}$ and the probability that it is smaller than that of S2 (or S14) is also significant, i.e., $\mathrel{\mathpalette\oversim>} 55\%$ (or $38\%$). The existence of such a star will provide a probe for testing the GR effects in the vicinity of an MBH. Future observations by the next generation telescopes, such as, TMT or E-ELT, will be able to investigate the existence of such a star, and provide important constraints on the nature of the central MBH if such a star is detected. The number of the ejected unbound companions of the ``GC S-stars'' (see the definition of the ``GC S-stars'' at the end of Section~\ref{sec:overview}) is roughly $\sim 20$-$40$ and the majority of these ejected stars are located within a distance of $\sim 20$~kpc from the GC. The number of these ejected companions is substantially larger than the number of observed GC S-stars mainly because the observed GC S-stars are only a fraction of the ``GC S-stars'' and the rest of the ``GC S-stars'' were tidally disrupted and do not survive today (see Table~\ref{tab:t4}). Their heliocentric radial velocities in the Galactic rest frame range from $\sim -500{\rm km\,s^{-1}}$ to $\sim 1500{\rm km\,s^{-1}}$ and their proper motions in the heliocentric rest frame can be as large as $\sim20{\rm mas\,yr^{-1}}$. These high-mass ejected stars are bright enough to be detected at a distance less than a few ten kpc and their proper motions are also large enough to be measured by future telescopes, such as {\it Gaia}. The majority of the ejected companions of the GC S-stars lie outside the area surveyed by \citet{Brown09a} for our observers located at the Sun. \acknowledgements We thank the referee and the scientific editor, Eric Feigelson, for helpful comments and suggestions. We thank Warren Brown for useful comments on the paper and are grateful to Ann-Marie Madigan for helpful communications on the ARMA model for the dynamical evolution of the captured stars. This work was supported in part by the National Natural Science Foundation of China under nos. 10973001 and 10973017, and the BaiRen program from the National Astronomical Observatories, Chinese Academy of Sciences. {\bf \it Note added in proof.} After the submission of this paper, the following two new observational results have been reported, which are relevant to this work. (1) Meyer et al. (2012, Sci., 338, 84) discovered a faint star, S0-102, which is orbiting the MBH in the GC with shortest-known-period (11.5 yr). The existence of such a star is consistent with our predictions shown in Figure 7. (2) Lu et al. (2013, ApJ, 764, 155) estimate the initial mass function for stellar populations in the central 0.5 pc of the Galaxy and find it is top-heavy with a slope of $-1.7\pm 0.2$, which is consistent with the requirement by our model to re-produce the number ratio of HVSs to GC S-stars (see Table 4 and Section 5.1).
\section{Introduction}\label{aba:sec1} To understand the formation, evolution, and present-day properties of the Universe we should study the properties of structures forming the cosmic web - superclusters and galaxy groups and clusters in them together. We analyse the properties of clusters to determine substructures in them, study the relations between the multimodality of rich clusters from the SDSS DR8 and the environment where they reside, and compare the properties of clusters in superclusters of different morphology. We use data from the 8th data release of the Sloan Digital Sky Survey, in total 576493 galaxies from MAIN sample. Groups of galaxies are found with the Friends-of-Friends method, see \cite{2012A&A...540A.106T} for details. Data about 109 clusters with at least 50 member galaxies have been used. We apply the luminosity density field to determine superclusters of galaxies \cite{2012A&A...540A.106T, 2012A&A...539A..80L}, and find host superclusters for all clusters. \section{Methods and results} We apply several 3D, 2D, and 1D tests to analyse the presence of substructure in clusters, the location of their main galaxies, and the velocity distribution of galaxies in clusters, and find that more than 80\% of clusters from our sample demonstrate signs of multimodality (see also \cite{2012A&A...540A.123E}). The main galaxies of clusters are typically located near the center of one of the components in the cluster, but not always in the central component. In Figure~\ref{fig:grGnonG} we show the results of the two 3D test: the 3D normal mixture modelling and the Dressler-Shectman (DS) test. \begin{figure} \begin{center} \psfig{file=MG13_ME_f1.eps, width=4in} \caption{ The cluster 34726 (Abell cluster A2028). From left to right: the DS test bubble plot (symbol sizes are proportional to the probability to have substructure), and R.A. vs. Dec., and R.A. vs. velocity ($10^{2}km/sec$) plots; the symbols show different components as found with normal mixture modelling. The star shows the location of the main galaxy. } \label{fig:grGnonG} \end{center} \end{figure} We determined morphological types of all superclusters hosting rich clusters with Minkowski functionals. Using first three Minkowski functionals we defined shapefinders which describe the planarity (shapefinder $K_1$) and filamentarity (shapefinder $K_2$) of superclusters. The maximum value of the fourth Minkowski functional $V_3$ (the clumpiness) characterises the inner structure of the superclusters \cite{2007A&A...476..697E}. Superclusters show two main morphological types: spiders and filaments. Figure~\ref{fig:fil} shows the fourth Minkowski functional $V_3$ vs. mass fraction $mf$ and the shapefinders $K_1$ and $K_2$ (morphological signature) for a supercluster of filament morphology, SCl~027, and of spider morphology, SCl~019 \cite{2007A&A...476..697E, 2011ApJ...736...51E}. \begin{figure} \begin{center} \psfig{file=MG13_ME_f2a.eps, width=4in} \psfig{file=MG13_ME_f2b.eps, width=4in} \caption{ Sky distribution of galaxies (left panel), the fourth Minkowski functional $V_3$ (middle panel) and the shapefinder's $K_1$-$K_2$ plane (right panel) for a supercluster of filament morphology, SCl~027 (upper row), and of spider morphology, SCl~019 (lower row), two rich superclusters in the Sloan Great Wall. Black filled circles denote galaxies in clusters with at least 50 member galaxies, grey dots denote other galaxies. Mass fraction increases anti-clockwise along the $K_1$-$K_2$ curve. } \label{fig:fil} \end{center} \end{figure} \begin{figure} \begin{center} \psfig{file=MG13_ME_f3.eps, width=4in} \caption{ Cumulative distributions of the numbers of components in clusters, $N_{\mathrm{comp}}$, peculiar velocities of cluster main galaxies, $V_{\mathrm{pec}}$ {\bf (in $km~s^{-1}$)}, and p-value of the DS test, $p_{\mathrm{\Delta}}$ for clusters in superclusters of filament morphology (F, black solid line), of spider morphology (S, grey dashed line), and for isolated clusters (I, thin grey solid line). } \label{fig:cummorf} \end{center} \end{figure} Clusters in superclusters of spider morphology have higher probabilities to have substructure and larger peculiar velocities of their main galaxies than clusters in filament-type superclusters, being therefore dynamically younger (Fig.~\ref{fig:cummorf}). Superclusters of spider morphology have richer inner structure than superclusters of filament morphology with large number of filaments between clusters in them. This may lead to the differences noted in this study \cite{2012A&A...542A..36E}. We note that recently Costa-Duarte et al. \cite{2012arXiv1210.0455C} found that the galaxy populations in superclusters of filament and pancake type are similar. This shows that the problem is not yet solved. Future studies of the galaxies and groups in superclusters are needed to understand the role of the large scale environment in shaping the properties of galaxies and their systems. \section*{Acknowledgments} I thank my coauthors Jaan Vennik, Elmo Tempel, Enn Saar, Pasi Nurmi, Antti Ahvensalmi and others for a fruitful collaboration. The present study was supported by the Estonian Science Foundation grants No. 8005, 7765, 9428, and MJD272, by the Estonian Ministry for Education and Science research project SF0060067s08, and by the European Structural Funds grant for the Centre of Excellence "Dark Matter in (Astro)particle Physics and Cosmology" TK120. We are pleased to thank the SDSS Team for the publicly available data releases. \bibliographystyle{ws-procs975x65}
\section{Introduction\label{section introduction}} \textquotedblleft ISLAND, in physical geography, a term generally definable as a piece of land surrounded by water.\textquotedblright\ (Encyclop\ae dia Britannica, Eleventh Edition, Volume XIV, Cambridge University Press 1910.) Mathematical models of this definition were introduced and studied by several authors. These investigations utilized tools from different areas of mathematics, e.g. combinatorics, coding theory, lattice theory, analysis, fuzzy mathematics. Our goal is to provide a general setting that unifies these approaches. This general framework encompasses prime implicants of Boolean functions and concepts of a formal context as special cases, and it has close connections to graph theory and to proximity spaces. The notion of an island as a mathematical concept occurred first in Cz\'{e}dli \cite{czedli}, where a rectangular board was considered with a real number assigned to each cell of the board, representing the height of that cell. A set $S$ of cells forming a rectangle is called an \it island, \rm if the minimum height of $S$ is greater then the height of any cell around the perimeter of $S$, since in this case $S$ can become a piece of land surrounded by water after a flood producing an appropriate water level. The motivation to investigate such islands comes from Foldes and Singhi \cite{FS}, where islands on a $1\times n$ board (so-called full segments) played a key role in characterizing maximal instantaneous codes. The main result of \cite{czedli} is that the maximum number of islands on an $m\times n$ board is $\left\lfloor \left( mn+m+n-1\right) /2\right\rfloor $. However, the size of a system of islands (i.e., the collection of all islands appearing for given heights) that is maximal with respect to inclusion (not with respect to cardinality) can be as low as $m+n-1$ \cite{L}. Another important observation of \cite{czedli} is that any two islands are either comparable (i.e. one is contained in the other) or disjoint; moreover, disjoint islands cannot be too close to each other (i.e. they cannot have neighboring cells). It was also shown in \cite{czedli} that these properties actually characterize systems of islands. We refer to such a result as a \textquotedblleft dry\textquotedblright\ characterization, since it describes systems of islands in terms of intrinsic conditions, without referring to heights and water levels. The above mentioned paper \cite{czedli} of G\'{a}bor Cz\'{e}dli was a starting point for many investigations exploring several variations and various aspects of islands. Square islands on a rectangular board have been considered in \cite{HHNS,L3}, and islands have been studied also on cylindrical and toroidal boards \cite{BHH}, on triangular boards \cite{HNP,L2}, on higher dimensional rectangular boards \cite{P} as well as in a continuous setting \cite{LP,PPPSz}% . If we allow only a given finite subset of the reals as possible heights, then the problem of determining the maximum number of islands becomes considerably more difficult; see, e.g. \cite{HTM,HST2, MM}. Islands also appear naturally as cuts of lattice-valued functions \cite{HST}; furthermore, order-theoretic properties of systems of islands proved to be of interest on their own, and they have been investigated in lattices and partially ordered sets \cite{CzHaSch,CzSch,HR}. The notion of an island is an elementary combinatorial concept, yet it leads immediately to open problems, therefore it is a suitable topic to introduce students to mathematical research \cite{MV}. In this paper we introduce a general framework for islands that subsumes all of the earlier studied concepts of islands on finite boards. We will axiomatize those situations where islands have the \textquotedblleft comparable or disjoint\textquotedblright\ property mentioned above, and we will also present dry characterizations of systems of islands. \section{Definitions and examples} Our landscape is given by a nonempty base set $U$, and a function $h\colon U\rightarrow\mathbb{R}$ that assigns to each point $u\in U$ its height $h\left( u\right) $. If the minimum height $\min h\left( S\right) :=\min\left\{ h\left( u\right) \colon u\in S\right\} $ of a set $S\subseteq U$ is greater than the height of its surroundings, then $S$ can become an island if the water level is just below $\min h\left( S\right) $. To make this more precise, let us fix two families of sets $\mathcal{C}% ,\mathcal{K}\subseteq\mathcal{P}\left( U\right) $, where $\mathcal{P}\left( U\right) $ denotes the power set of $U$. We do not allow islands of arbitrary \textquotedblleft shapes\textquotedblright: only sets belonging to $\mathcal{C}$ are considered as candidates for being islands, and the members of $\mathcal{K}$ describe the \textquotedblleft surroundings\textquotedblright% \ of these sets. \begin{definition} An \emph{island domain} is a pair $\left( \mathcal{C},\mathcal{K}\right) $, where $\mathcal{C}\subseteq\mathcal{K}\subseteq\mathcal{P}\left( U\right) $ for some nonempty finite set $U$ such that $U\in\mathcal{C}$. By a \emph{height function} we mean a map $h\colon U\rightarrow\mathbb{R}$. \end{definition} Throughout the paper we will always implicitly assume that $\left( \mathcal{C}% ,\mathcal{K}\right) $ is an island domain. We denote the cover relation of the poset $\left( \mathcal{K},\subseteq\right) $ by $\prec$, and we write $K_{1}\preceq K_{2}$ if $K_{1}\prec K_{2}$ or $K_{1}=K_{2}$. \begin{definition} Let $\left( \mathcal{C},\mathcal{K}\right) $ be an island domain, let $h\colon U\rightarrow\mathbb{R}$ be a height function and let $S\in \mathcal{C}$ be a nonempty set.% \renewcommand{\theenumi}{(\roman{enumi})} \renewcommand{\labelenumi}{\theenumi}% \begin{enumerate} \item \label{def island (i)}We say that $S$ is a\emph{ pre-island} with respect to the triple $\left( \mathcal{C},\mathcal{K},h\right) $, if every $K\in\mathcal{K}$ with $S\prec K$ satisfies% \[ \min h\left( K\right) <\min h\left( S\right) . \] \item \label{def island (ii)}We say that $S$ is an \emph{island} with respect to the triple $\left( \mathcal{C},\mathcal{K},h\right) $, if every $K\in\mathcal{K}$ with $S\prec K$ satisfies \[ h\left( u\right) <\min h\left( S\right) \text{ for all }u\in K\setminus S. \] \end{enumerate} The \emph{system of (pre-)islands corresponding to }$\left( \mathcal{C}% ,\mathcal{K},h\right) $ is the set% \[ \left\{ S\in\mathcal{C}\setminus\left\{ \emptyset\right\} \colon S\text{ is a (pre-)island w.r.t. }\left( \mathcal{C},\mathcal{K},h\right) \right\} . \] By a \emph{system of (pre-)islands corresponding to }$\left( \mathcal{C}% ,\mathcal{K}\right) $ we mean a set $\mathcal{S}\subseteq\mathcal{C}$ such that there is a height function $h\colon U\rightarrow\mathbb{R}$ so that the system of (pre-)islands corresponding to $\left( \mathcal{C},\mathcal{K}% ,h\right) $ is $\mathcal{S}$. \end{definition} \begin{remark} \label{remark basic}Let us make some simple observations concerning the above definition.% \renewcommand{\theenumi}{(\alph{enumi})} \renewcommand{\labelenumi}{\theenumi}% \begin{enumerate} \item Every nonempty set $S$ in $\mathcal{C}$ is in fact an island for some height function $h.$ \item If $S$ is an island with respect to $\left( \mathcal{C},\mathcal{K}% ,h\right) $, then $S$ is also a pre-island with respect to $\left( \mathcal{C},\mathcal{K},h\right) $. The converse is not true in general; however, if for every nonempty\emph{ }$C\in\mathcal{C}$ and $K\in\mathcal{K}$ with\emph{ }$C\prec K$ we have $\left\vert K\setminus C\right\vert =1$, then the two notions coincide. \item The set $U$ is always a (pre-)island. If $S$ is a (pre-)island that is different from $U$, then we say that $S$ is a \emph{proper (pre-)island}. \item If $S$ is a pre-island with respect to $\left( \mathcal{C}% ,\mathcal{K},h\right) $, then the inequality $\min h\left( K\right) <\min h\left( S\right) $ of \ref{def island (i)} holds for all $K\in\mathcal{K}$ with $S\subset K$ (not just for covers of $S)$. \item Let $\mathcal{C}\subseteq\mathcal{K}^{\prime}\subseteq\mathcal{K}$. It is easy to see that any $\mathcal{S}\in\mathcal{C}$ which is a pre-island with respect to the triple $\left( \mathcal{C},\mathcal{K},h\right) $ is also a pre-island with respect to $\left( \mathcal{C},\mathcal{K}^{\prime},h\right) $. \item The numerical values of the height function $h$ are not important; only the partial ordering that $h$ establishes on $U$ is relevant. In particular, one could assume without loss of generality that the range of $h$ is contained in the set $\left\{ 0,1,\ldots,\left\vert U\right\vert -1\right\} $. \end{enumerate} \end{remark} Many of the previously studied island concepts can be interpreted in terms of graphs as follows. \begin{example} \label{example graphs}Let $G=\left( U,E\right) $ be a connected simple graph with vertex set $U$ and edge set $E$; let $\mathcal{K}$ consist of the connected subsets of $U$, and let $\mathcal{C}\subseteq\mathcal{K}$ such that $U\in\mathcal{C}$. In this case the second item of Remark~\ref{remark basic} applies, hence pre-islands and islands are the same. Let us assume that $G$ is connected, and let $\mathcal{C}$ consist of the connected convex sets of vertices. (A set is called convex if it contains all shortest paths between any two of its vertices.) If $G$ is a path, then the islands are exactly the full segments considered in \cite{FS}, and if $G$ is a square grid (the product of two paths), then we obtain the rectangular islands of \cite{czedli}. Square islands on a rectangular board \cite{HHNS,L3}, islands on cylindrical and toroidal boards \cite{BHH}, on triangular boards \cite{HNP,L2} and on higher dimensional rectangular boards \cite{P} also fit into this setting. \end{example} Surprisingly, formal concepts and prime implicants are also pre-islands in disguise. \begin{example} \label{example context}Let $A_{1},\ldots,A_{n}$ be nonempty sets, and let $\mathcal{I}\subseteq A_{1}\times\cdots\times A_{n}$. Let us define% \begin{align*} U & =A_{1}\times\cdots\times A_{n},\\ \mathcal{K} & =\left\{ B_{1}\times\cdots\times B_{n}\colon\emptyset\neq B_{i}\subseteq A_{i},~1\leq i\leq n\right\} \\ \mathcal{C} & =\left\{ C\in\mathcal{K}\colon C\subseteq\mathcal{I}\right\} \cup\{U\}, \end{align*} and let $h\colon U\longrightarrow\{0,1\}$ be the height function given by% \[ h\left( a_{1},\ldots,a_{n}\right) :=\left\{ \!\!% \begin{array} [c]{rl}% 1\text{,} & \text{if }\left( a_{1},\ldots,a_{n}\right) \in\mathcal{I};\\ 0\text{,} & \text{if }\left( a_{1},\ldots,a_{n}\right) \in U\setminus \mathcal{I}; \end{array} \right. \text{ for all }\left( a_{1},\ldots,a_{n}\right) \in U. \] It is easy to see that the pre-islands corresponding to the triple $\left( \mathcal{C},\mathcal{K},h\right) $ are exactly $U$ and the maximal elements of the poset $\left( \mathcal{C}\setminus\left\{ U\right\} ,\subseteq \right) $. Now let $\left( G,M,\mathcal{I}\right) $, $\mathcal{I}\subseteq G\times M$ be a formal context, and let us apply the above construction with $A_{1}=G$, $A_{2}=M$ and $U=A_{1}\times A_{2}$. Then the pre-islands are $U$ and the concepts of the context $(G,M,\mathcal{I)}$ with nonempty extent and intent \cite{GW}. Further, consider the case $A_{1}=\cdots=A_{n}=\{0,1\}$. Then the height function $h$ is an $n$-ary Boolean function, and it is not hard to check that the pre-islands corresponding to $\left( \mathcal{C},\mathcal{K},h\right) $ are $U$ and the prime implicants of $h$ \cite{CrHa}. \end{example} \begin{remark} \label{rem refinement}For any given island domain $\left( \mathcal{C}% ,\mathcal{K}\right) $, maximal families of (pre-)islands are realized by injective height functions. To see this, let us assume that $h$ is a non-injective height function, i.e. there exists a number $z$ in the range of $h$ such that $h^{-1}\left( z\right) =\left\{ s_{1},\ldots,s_{m}\right\} $ with $m\geq2$. The following \textquotedblleft refinement\textquotedblright% \ procedure constructs another height function $g$ so that every (pre-)island corresponding to $\left( \mathcal{C},\mathcal{K},h\right) $ is also a (pre-)island with respect to $\left( \mathcal{C},\mathcal{K},g\right) $. Let $y$ be the largest value of $h$ below $z$ (or $z-1$ if $z$ is the minimum value of the range of $h$), and let $w$ be the smallest value of $h$ above $z$ (or $z+1$ if $z$ is the maximum value of the range of $h$). For any $u\in U$, we define $g\left( u\right) $ by% \[ g\left( u\right) =\left\{ \!\!\renewcommand{\arraystretch}{1.8}% \begin{array} [c]{rl}% y+i\dfrac{w-y}{m+1}, & \text{if }u=s_{i};\\ h\left( u\right) , & \text{if }h\left( u\right) \neq z. \end{array} \right. \] By repeatedly applying this procedure we obtain an injective height function without losing any pre-islands. Note that injective height functions correspond to linear orderings of $U$ (cf. the last observation of Remark~\ref{remark basic}). \end{remark} \begin{example} \label{example projective plane}Let $U$ be a finite projective plane of order $p$, thus $U$ has $m:=p^{2}+p+1$ points. Let $\mathcal{C}=\mathcal{K}$ consist of the whole plane, the lines, the points and the empty set. Then the greatest possible number of pre-islands is $p^{2}+2=m-p+1$. Indeed, as explained in Remark~\ref{rem refinement}, the largest systems of pre-islands emerge with respect to linear orderings of $U$. So let us consider a linear order on $U$, and let $\mathbf{0}$ and $\mathbf{1}$ denote the smallest and largest elements of $U$, respectively. In other words, we have $h\left( \mathbf{0}\right) <h(x)<h\left( \mathbf{1}\right) $ for all $x\in U\setminus\{\mathbf{0}% ,\mathbf{1}\}$. Clearly, a line is a pre-island iff it does not contain $\mathbf{0}$, and there are $m-p-1$ such lines. The only other pre-islands are the point $\mathbf{1}$ and the entire plane, hence we obtain $m-p-1+2=m-p+1$ pre-islands. \end{example} It has been observed in \cite{czedli, HNP,HHNS} that any two islands on a square or triangular grid with respect to a given height function are either comparable or disjoint. This property is formalized in the following definition, which was introduced in \cite{CzHaSch}. \begin{definition} A family $\mathcal{H}$ of subsets of $U$ is $\operatorname{CD}$% \emph{-independent} if any two members of $\mathcal{H}$ are either comparable or disjoint, i.e. for all $A,B\in\mathcal{H}$ at least one of $A\subseteq B$, $B\subseteq A$ or $A\cap B=\emptyset$ holds. \end{definition} Note that $\operatorname{CD}$-independence is also known as laminarity \cite{LP,PPPSz}. In general, the properties of $\operatorname{CD}% $-independence and being a system of pre-islands are independent from each other, as the following example shows. \begin{example} \label{example not CD}Let $U=\left\{ a,b,c,d,e\right\} $ and $\mathcal{K}% =\mathcal{C}=\left\{ \left\{ a,b\right\} ,\left\{ a,c\right\} ,\left\{ b,d\right\} ,\left\{ c,d\right\} ,U\right\} $. Let us define a height function $h$ on $U$ by $h\left( a\right) =h\left( b\right) =h\left( c\right) =h\left( d\right) =1$, $h\left( e\right) =0$. It is easy to verify that every element of $\mathcal{C}$ is a pre-island with respect to this height function, but $\mathcal{C}$ is not $\operatorname{CD}% $-independent. On the other hand, consider the $\operatorname{CD}$-independent family $\mathcal{H}=\left\{ \left\{ a,b\right\} ,\left\{ c,d\right\} ,U\right\} $. We claim that $\mathcal{H}$ is not a system of pre-islands. To see this, assume that $h$ is a height function such that the system of pre-islands corresponding to $\left( \mathcal{C},\mathcal{K},h\right) $ is $\mathcal{H}$. Let us write out the definition of a pre-island for $S=\left\{ a,b\right\} $ and $S=\left\{ c,d\right\} $ with $K=U$:% \begin{align*} \min\left( h\left( a\right) ,h\left( b\right) \right) & >\min h\left( U\right) ;\\ \min\left( h\left( c\right) ,h\left( d\right) \right) & >\min h\left( U\right) . \end{align*} Taking the minimum of these two inequalities, we obtain% \[ \min\left( h\left( a\right) ,h\left( b\right) ,h\left( c\right) ,h\left( d\right) \right) >\min h\left( U\right) . \] This immediately implies that $\min\left( h\left( a\right) ,h\left( c\right) \right) >\min h\left( U\right) $. Since the only element of $\mathcal{K}$ properly containing $\left\{ a,c\right\} $ is $U$, we can conclude that $\left\{ a,c\right\} $ is also a pre-island with respect to $h$, although $\left\{ a,c\right\} \notin\mathcal{H}$. \end{example} As $\operatorname{CD}$-independence is a natural and desirable property of islands that was crucial in previous investigations, we will mainly focus on island domains $\left( \mathcal{C},\mathcal{K}\right) $ whose systems of pre-islands are $\operatorname{CD}$-independent. We characterize such island domains in Theorem~\ref{thm ID <==> (SZ==>CD)}, and we refer to them as \emph{connective island domains} (see Definition~\ref{def ID}). The most fundamental questions concerning pre-islands are the following: Given an island domain $\left( \mathcal{C},\mathcal{K}\right) $ and a family $\mathcal{H}\subseteq\mathcal{C}$, how can we decide if there is a height function $h$ such that $\mathcal{H}$ is the system of pre-islands corresponding to $\left( \mathcal{C},\mathcal{K},h\right) $? How can we find such a height function (if there is one)? Concerning the first question, we give a dry\ characterization of systems of pre-islands corresponding to connective island domains in Theorem~\ref{thm ID ==> (HA<==>SZ)}, and in Corollary~\ref{cor PD ==> (distant<==>strSZ)} we characterize systems of islands corresponding to so-called \emph{proximity domains} (see Definition~\ref{def proximity domain}). These results generalize earlier dry characterizations (see, e.g. \cite{czedli, HNP,HHNS}), since an island domain $\left( \mathcal{C},\mathcal{K}\right) $ corresponding to a graph (cf. Example~\ref{example graphs}) is always a connective island domain and also a proximity domain. Concerning the second question, we give a canonical construction for a height function (Definition~\ref{def canonical height function}), and we prove in Sections~\ref{section CD and ID} and \ref{section strict} that this height function works for pre-islands in connective island domains and for islands in proximity domains. \section{Pre-islands and admissible systems\label{section admissible}} In this section we present a condition that is necessary for being a system of pre-islands, which will play a key role in later sections. Although this necessary condition is not sufficient in general, we will use it to obtain a characterization of \emph{maximal} systems of pre-islands. \begin{definition} \label{def admissible}Let $\mathcal{H}\subseteq\mathcal{C}\setminus\left\{ \emptyset\right\} $ be a family of sets such that $U\in\mathcal{H}$. We say that $\mathcal{H}$ is \emph{admissible (with respect to }$\left( \mathcal{C},\mathcal{K}\right) $\emph{)}, if for every nonempty antichain $\mathcal{A}\subseteq\mathcal{H}$,% \begin{equation} \exists H\in\mathcal{A}\text{ such that }\forall K\in\mathcal{K}:~H\subset K\implies K\nsubseteq\bigcup\,\mathcal{A}.\label{eq admissible}% \end{equation} \end{definition} \begin{remark} \label{rem admissible not just for antichains}Let us note that if $\mathcal{H}$ is admissible, then (\ref{eq admissible}) holds for \emph{all} nonempty $\mathcal{A}\subseteq\mathcal{H}$ (not just for antichains). Indeed, if $\mathcal{M}$ denotes the set of maximal members of $\mathcal{A}$, then $\mathcal{M}$ is an antichain. Thus the admissibility of $\mathcal{H}$ implies that there is $H\in\mathcal{M}\subseteq\mathcal{A}$ such that for all $K\in\mathcal{K}$ with $H\subset K$ we have $K\nsubseteq\bigcup\,\mathcal{M}% =\bigcup\,\mathcal{A}$. \end{remark} Obviously, any subfamily of an admissible family is also admissible, provided that it contains $U$. As we shall see later, in some important special cases a stronger version of admissibility holds, where the existential quantifier is replaced by a universal quantifier in (\ref{eq admissible}): for every nonempty antichain $\mathcal{A}\subseteq\mathcal{H}$,% \begin{equation} \forall H\in\mathcal{A~}\forall K\in\mathcal{K}:~H\subset K\implies K\nsubseteq\bigcup\,\mathcal{A}. \label{eq stronger admissible}% \end{equation} \begin{proposition} \label{prop SZ==>HA}Every system of pre-islands is admissible. \end{proposition} \begin{proof} Let $h\colon U\rightarrow\mathbb{R}$ be a height function and let $\mathcal{S}$ be the system of pre-islands corresponding to $\left( \mathcal{C},\mathcal{K},h\right) $. Clearly, we have $\emptyset \notin\mathcal{S}$ and $U\in\mathcal{S}$. Let us assume for contradiction that there exists an antichain $\mathcal{A}=\left\{ S_{i}:i\in I\right\} \subseteq\mathcal{S}$ such that (\ref{eq admissible}) does not hold. Then for every $i\in I$ there exists $K_{i}\in\mathcal{K}$ such that $S_{i}\subset K_{i}$ and $K_{i}\subseteq\bigcup_{i\in I}S_{i}$. Since $S_{i}$ is a pre-island, we have% \[ \min h\left( S_{i}\right) >\min h\left( K_{i}\right) \geq\min h% \Bigl(% \bigcup_{i\in I}S_{i}% \Bigr)% \] for all $i\in I$. Taking the minimum of these inequalities we arrive at the contradiction% \[ \min\left\{ \min h\left( S_{i}\right) \mid i\in I\right\} >\min h% \Bigl(% \bigcup_{i\in I}S_{i}% \Bigr)% .% \qedhere \] \end{proof} The converse of Proposition~\ref{prop SZ==>HA} is not true in general: it is straightforward to verify that the family $\mathcal{H}$ considered in Example~\ref{example not CD} is admissible, but, as we have seen, it is not a system of pre-islands. However, we will prove in Proposition~\ref{prop HA==>SZ resze} that for every admissible family $\mathcal{H}$, there exists a height function such that the corresponding system of pre-islands contains $\mathcal{H}$. First we give the construction of this height function, and we illustrate it with some examples. \begin{definition} \label{def canonical height function}Let $\mathcal{H}\subseteq\mathcal{C}$ be an admissible family of sets. We define subfamilies $\mathcal{H}^{\left( i\right) }\subseteq\mathcal{H}~\left( i=0,1,2,\ldots\right) $ recursively as follows. Let $\mathcal{H}^{\left( 0\right) }=\left\{ U\right\} $. For $i>0$, if $\mathcal{H}\neq\mathcal{H}^{\left( 0\right) }\cup\cdots \cup\mathcal{H}^{\left( i-1\right) }$, then let $\mathcal{H}^{\left( i\right) }$ consist of all those sets $H\in\mathcal{H}\setminus (\mathcal{H}^{\left( 0\right) }\cup\cdots\cup\mathcal{H}^{\left( i-1\right) })$ that have the following property:% \begin{equation} \forall K\in\mathcal{K}:~H\subset K\implies K\nsubseteq\bigcup\,% \bigl(% \mathcal{H}\setminus(\mathcal{H}^{\left( 0\right) }\cup\cdots\cup \mathcal{H}^{\left( i-1\right) })% \bigr)% . \label{eq def canonical height function: set}% \end{equation} Since $\mathcal{H}$ is finite and admissible, after finitely many steps we obtain a partition $\mathcal{H}=\mathcal{H}^{\left( 0\right) }\cup\cdots \cup\mathcal{H}^{\left( r\right) }$ (cf. Remark~\ref{rem admissible not just for antichains}). The \emph{canonical height function corresponding to }$\mathcal{H}$ is the function $h_{\mathcal{H}}\colon U\rightarrow\mathbb{N}$ defined by% \begin{equation} h_{\mathcal{H}}\left( x\right) :=\max\left\{ i\in\left\{ 1,\ldots ,r\right\} :x\in\bigcup\,\mathcal{H}^{\left( i\right) }\right\} \text{ for all }x\in U. \label{eq def canonical height function: height}% \end{equation} \end{definition} Observe that every $\mathcal{H}^{\left( i\right) }$ consists of \emph{some} of the maximal members of $\mathcal{H}\setminus(\mathcal{H}^{\left( 0\right) }\cup\cdots\cup\mathcal{H}^{\left( i-1\right) })=\mathcal{H}^{\left( i\right) }\cup\cdots\cup\mathcal{H}^{\left( r\right) }$. However, if $\mathcal{H}$ satisfies (\ref{eq stronger admissible}) for all antichains $\mathcal{A}\subseteq\mathcal{H}$, then the word \textquotedblleft% \emph{some}\textquotedblright\ can be replaced by \textquotedblleft% \emph{all}\textquotedblright\ in the previous sentence, and in this case $h_{\mathcal{H}}$ can be computed just from $\mathcal{H}$ itself, without making reference to $\mathcal{K}$. To illustrate this, let us consider a $\operatorname{CD}$-independent family $\mathcal{H}$. Clearly, for every $u\in U$, the set of members of $\mathcal{H}$ containing $u$ is a finite chain. The \emph{standard height function} of $\mathcal{H}$ assigns to each element $u$ the length of this chain, i.e. one less than the number of members of $\mathcal{H}$ that contain $u$. (Note that the definition of a standard height function in \cite{HST2} differs slightly from ours.) It is easy to see that if $\mathcal{H}$ satisfies (\ref{eq stronger admissible}), then the canonical height function of $h$ coincides with the standard height function. However, in general the two functions might be different. Figure~\ref{fig standardeskanonikus} represents the standard and the canonical height functions for the same $\operatorname{CD}$-independent family, with greater heights indicated by darker colors. We can see from Figure~\ref{fig kanonikus} that only two of the four maximal members of $\mathcal{H}\setminus\left\{ U\right\} $ belong to $\mathcal{H}^{\left( 1\right) }$, thus (\ref{eq stronger admissible}) fails here. (In order to make the picture comprehensible, only members of $\mathcal{C}$ are shown, although $\mathcal{K}$ is also needed to determine $h_{\mathcal{H}}$ (Figure~\ref{fig kanonikus}). On the other hand, the standard height function (Figure~\ref{fig standard}) can be read directly from the figure.) \begin{figure}[h] \centering \subfloat[Standard height function] { \includegraphics[width=0.4\textwidth]{standardb.pdf} \label{fig standard}} \qquad\subfloat[Canonical height function]{ \includegraphics[width=0.4\textwidth]{kanonikusb.pdf} \label{fig kanonikus}}\caption{A $\operatorname{CD}$-independent family with two different height functions}% \label{fig standardeskanonikus}% \end{figure} The next example shows that there exist $\operatorname{CD}$-independent systems of pre-islands for which the standard height function is not the right choice. However, in Section~\ref{section strict} we will see that for a wide class of island domains, including those corresponding to graphs (cf. Example~\ref{example graphs}), the standard height function is always appropriate. \begin{example} \label{example not standard}Let $U=\left\{ a,b,c,d\right\} $, $\mathcal{C}% =\left\{ A,B,U\right\} $ and $\mathcal{K}=\left\{ A,B,U,K\right\} $, where $A=\left\{ a\right\} $, $B=\left\{ b,c\right\} $ and $K=\left\{ a,c\right\} $. Then the family $\mathcal{H}=\left\{ A,B,U\right\} $ is admissible; the corresponding partition is $\mathcal{H}^{\left( 0\right) }=\left\{ U\right\} $, $\mathcal{H}^{\left( 1\right) }=\left\{ B\right\} $, $\mathcal{H}^{\left( 2\right) }=\left\{ A\right\} $, and the canonical height function is given by $h_{\mathcal{H}}\left( a\right) =2$, $h_{\mathcal{H}}\left( b\right) =h_{\mathcal{H}}\left( c\right) =1$, $h_{\mathcal{H}}\left( d\right) =0$. It is straightforward to verify that $\mathcal{H}$ is the system of pre-islands corresponding to $\left( \mathcal{C},\mathcal{K},h_{\mathcal{H}}\right) $. However, the standard height function assigns the value $1$ to $a$, and thus $A$ is not a pre-island with respect to the standard height function of $\mathcal{H}$. \end{example} \begin{proposition} \label{prop HA==>SZ resze}If $\mathcal{H\subseteq C}$ is an admissible family of sets and $h_{\mathcal{H}}$ is the corresponding canonical height function, then every member of $\mathcal{H}$ is a pre-island with respect to $\left( \mathcal{C},\mathcal{K},h_{\mathcal{H}}\right) $. \end{proposition} \begin{proof} Let $\mathcal{H\subseteq C}$ be admissible, and let us consider the partition $\mathcal{H}=\mathcal{H}^{\left( 0\right) }\cup\cdots\cup\mathcal{H}% ^{\left( r\right) }$ given in Definition~\ref{def canonical height function}% . For each $H\in\mathcal{H}$, there is a unique $i\in\left\{ 1,\ldots ,r\right\} $ such that $H\in\mathcal{H}^{\left( i\right) }$, and we have $\min h_{\mathcal{H}}\left( H\right) \geq i$ by (\ref{eq def canonical height function: height}). Using this observation it is straightforward to verify that $H$ is indeed a pre-island with respect to $\left( \mathcal{C},\mathcal{K},h_{\mathcal{H}}\right) $. \end{proof} As an immediate consequence of Propositions~\ref{prop SZ==>HA} and \ref{prop HA==>SZ resze} we have the following corollary. \begin{corollary} \label{cor maximal HA <==> maximal SZ}A subfamily of $\mathcal{C}$ is a maximal system of pre-islands if and only if it is a maximal admissible family. \end{corollary} We have seen in Example~\ref{example not CD} that it is possible that a subset of a system of pre-islands is not a system of pre-islands. The notion of admissibility allows us to describe those situations where this cannot happen. \begin{proposition} \label{prop subsystem}The following two conditions are equivalent for any island domain $\left( \mathcal{C},\mathcal{K}\right) $:% \renewcommand{\theenumi}{(\roman{enumi})} \renewcommand{\labelenumi}{\theenumi}% \begin{enumerate} \item \label{prop subsystem (i)}Any subset of a system of pre-islands corresponding to $\left( \mathcal{C},\mathcal{K}\right) $ that contains $U$ is also a system of pre-islands. \item \label{prop subsystem (ii)}The systems of pre-islands corresponding to $\left( \mathcal{C},\mathcal{K}\right) $ are exactly the admissible families. \end{enumerate} \end{proposition} \begin{proof} The implication \ref{prop subsystem (ii)}$\implies$\ref{prop subsystem (i)} follows from the simple observation that any subset of an admissible family containing $U$ is also admissible. Assume now that \ref{prop subsystem (i)} holds. In view of Proposition~\ref{prop SZ==>HA}, it suffices to prove that every admissible family is a system of pre-islands. Let $\mathcal{H}$ be an admissible family, then Proposition~\ref{prop HA==>SZ resze} yields a system of pre-islands containing $\mathcal{H}$. Using \ref{prop subsystem (i)} we can conclude that $\mathcal{H}$ is a system of pre-islands. \end{proof} \section{$\operatorname{CD}$-independence and connective island domains\label{section CD and ID}} As we have seen in Example~\ref{example not CD}, a system of pre-islands is not necessarily $\operatorname{CD}$-independent. In this section we present a condition that characterizes those island domains $\left( \mathcal{C}% ,\mathcal{K}\right) $ whose systems of pre-islands are $\operatorname{CD}% $-independent, and we will prove that admissibility is necessary and sufficient for being a systems of pre-islands in this case. \begin{definition} \label{def ID}An island domain $\left( \mathcal{C},\mathcal{K}\right) $ is a \emph{connective island domain} if% \begin{equation} \forall A,B\in\mathcal{C}:~\left( A\cap B\neq\emptyset\text{ and }B\nsubseteq A\right) \implies\exists K\in\mathcal{K}:A\subset K\subseteq A\cup B. \label{eq def ID}% \end{equation} \end{definition} \begin{remark} \label{rem ID symmetry}Observe that if $A\subset B$, then (\ref{eq def ID}) is satisfied with $K=B$. Thus it suffices to require (\ref{eq def ID}) for sets $A,B$ that are not comparable or disjoint. In this case, by switching the role of $A$ and $B$, we obtain that there is also a set $K^{\prime}\in\mathcal{K}$ such that $B\subset K^{\prime}\subseteq A\cup B$ (see Figure~\ref{fig islanddomain}). \end{remark} \begin{figure} [tb] \begin{center} \includegraphics[ natheight=4.002400in, natwidth=6.353600in, height=4.0958cm, width=6.4617cm ]% {islanddomainb.pdf}% \caption{Illustration to the definition of an island domain}% \label{fig islanddomain}% \end{center} \end{figure} \begin{remark} The terminology is motiveted by the intuition that the set $K$ in Definition~\ref{def ID} somehow connects $A$ and $B$. Let us note that if $\left( \mathcal{C},\mathcal{K}\right) $ corresponds to a graph, as in Example~\ref{example graphs}, then $\left( \mathcal{C},\mathcal{K}\right) $ is a connective island domain. Furthermore, it is not difficult to prove that if $\left( \mathcal{C},\mathcal{K}\right) $ is a connective island domain with $\mathcal{C}=\mathcal{K}$, then (\ref{eq def ID}) is equivalent to the fact that the union of two overlapping members of $\mathcal{K}$ belongs to $\mathcal{K}$ (see (\ref{eq KK}) in\ Section~\ref{section strict}), which is an important property of connected sets. \end{remark} We will prove that pre-islands corresponding to connective island domains are not only $\operatorname{CD}$-independent, but they also satisfy the following stronger independence condition, usually called $\operatorname*{CDW}% $-independence, which was introduced in \cite{CzSch}. \begin{definition} \label{def CDW-independence}A family $\mathcal{H}\subseteq\mathcal{P}\left( U\right) $ is \emph{weakly independent} (see \cite{CzHSch}) if \begin{equation} H\subseteq\bigcup_{i\in I}H_{i}\implies\exists i\in I:H\subseteq H_{i} \label{eq def CDW-independence}% \end{equation} holds for all $H\in\mathcal{H},H_{i}\in\mathcal{H}\left( i\in I\right) $. If $\mathcal{H}$ is both $\operatorname{CD}$-independent and weakly independent, then we say that $\mathcal{H}$ is $\operatorname*{CDW}$\emph{-independent}. \end{definition} \begin{remark} \label{rem CD maximal subsets}Let $\mathcal{H}\subseteq\mathcal{P}\left( U\right) $ be a $\operatorname{CD}$-independent family, and let $H\in\mathcal{H}$. Let $M_{1},\ldots,M_{m}$ be those elements of $\mathcal{H}$ that are properly contained in $H$ and are maximal with respect to this property. Then $M_{1},\ldots,M_{m}$ are pairwise disjoint, and $M_{1}% \cup\cdots\cup M_{m}\subseteq H$. Weak independence of $\mathcal{H}$ is equivalent to the fact that this latter containment is strict for every $H\in\mathcal{H}$. In particular, in the definition of weak independence it suffices to require (\ref{eq def CDW-independence}) for pairwise disjoint sets $H_{i}$. \end{remark} \begin{lemma} \label{lemma ID ==> (HA==>CDW)}If $\left( \mathcal{C},\mathcal{K}\right) $ is a connective island domain, then every admissible subfamily of $\mathcal{C}$ is $\operatorname*{CDW}$-independent. \end{lemma} \begin{proof} Let $\left( \mathcal{C},\mathcal{K}\right) $ be a connective island domain, and let $\mathcal{H}\subseteq\mathcal{C}$ be an admissible family. If $A,B\in\mathcal{H}$ are neither comparable nor disjoint, then (\ref{eq def ID}% ) and Remark~\ref{rem ID symmetry} show that $\mathcal{A}:=\left\{ A,B\right\} $ is an antichain for which (\ref{eq admissible}) does not hold (see Figure~\ref{fig islanddomain}). Thus $\mathcal{H}$ is $\operatorname{CD}$-independent. To prove that $\mathcal{H}$ is also $\operatorname*{CDW}$-independent, we apply Remark~\ref{rem CD maximal subsets}. Let us assume for contradiction that $M_{1}\cup\cdots\cup M_{m}=H$ for pairwise disjoint sets $M_{1}% ,\ldots,M_{m}\in\mathcal{H}\left( m\geq2\right) $ and $H\in\mathcal{H}$. Since $M_{i}\subset H\in\mathcal{K}$ and $H\subseteq M_{1}\cup\cdots\cup M_{m}$ for $i=1,\ldots,m$, we see that (\ref{eq admissible}) fails for the antichain $\mathcal{A}:=\left\{ M_{1},\ldots,M_{m}\right\} $, contradicting the admissibility of $\mathcal{H}$. \end{proof} As the next example shows, a $\operatorname*{CDW}$-independent family in a connective island domain is not necessarily admissible. \begin{example} \label{example CDNT but not HA}Let us consider the same sets $U$, $A$, $B$ and $K$ as in Example$~$\ref{example not standard}, and let $\mathcal{C}=\left\{ A,B,U\right\} $ and $\mathcal{K}=\left\{ A,B,U,K,L\right\} $, where $L=\left\{ a,b,c\right\} $. Then $\left( \mathcal{C},\mathcal{K}\right) $ is a connective island domain and $\left\{ A,B,U\right\} $ is $\operatorname*{CDW}$-independent, but it is not admissible (hence not a system of pre-islands). \end{example} \begin{theorem} \label{thm ID <==> (SZ==>CD)}The following three conditions are equivalent for any island domain $\left( \mathcal{C},\mathcal{K}\right) $:% \renewcommand{\theenumi}{(\roman{enumi})} \renewcommand{\labelenumi}{\theenumi}% \begin{enumerate} \item \label{thm ID <==> (SZ==>CD) (i)}$\left( \mathcal{C},\mathcal{K}% \right) $ is a connective island domain. \item \label{thm ID <==> (SZ==>CD) (ii)}Every system of pre-islands corresponding to $\left( \mathcal{C},\mathcal{K}\right) $ is $\operatorname{CD}$-independent. \item \label{thm ID <==> (SZ==>CD) (iii)}Every system of pre-islands corresponding to $\left( \mathcal{C},\mathcal{K}\right) $ is $\operatorname*{CDW}$-independent. \end{enumerate} \end{theorem} \begin{proof} It is obvious that \ref{thm ID <==> (SZ==>CD) (iii)}$\implies$% \ref{thm ID <==> (SZ==>CD) (ii)}. To prove that \ref{thm ID <==> (SZ==>CD) (ii)}$\implies$% \ref{thm ID <==> (SZ==>CD) (i)}, let us assume that $\left( \mathcal{C}% ,\mathcal{K}\right) $ is not a connective island domain. Then there exist $A,B\in\mathcal{C}$ that are not comparable or disjoint such that there is no $K\in\mathcal{K}$ with $A\subset K\subseteq A\cup B$. We define a height function $h\colon U\rightarrow\mathbb{N}$ as follows:% \[ h\left( x\right) :=\left\{ \!\!% \begin{array} [c]{rl}% 2, & \text{if }x\in B;\\ 1, & \text{if }x\in A\setminus B;\\ 0, & \text{if }x\notin A\cup B\text{.}% \end{array} \right. \] We claim that both $A$ and $B$ are pre-islands with respect to $\left( \mathcal{C},\mathcal{K},h\right) $. This is clear for $B$, as $\min h\left( K\right) \leq1$ for any proper superset $K$ of $B$. On the other hand, our assumption implies that for any $K\supset A$ we have $K\nsubseteq A\cup B$, hence $\min h\left( K\right) =0<\min h\left( A\right) =1$, thus $A$ is indeed a pre-island. Since $A$ and $B$ are not $\operatorname{CD}$, the system of pre-islands corresponding to $\left( \mathcal{C},\mathcal{K},h\right) $ is not $\operatorname{CD}$-independent. Finally, for the implication \ref{thm ID <==> (SZ==>CD) (i)}$\implies $\ref{thm ID <==> (SZ==>CD) (iii)}, assume that $\left( \mathcal{C}% ,\mathcal{K}\right) $ is a connective island domain and $\mathcal{S}$ is a system of pre-islands corresponding to $\left( \mathcal{C},\mathcal{K}% \right) $. By Proposition~\ref{prop SZ==>HA}, $\mathcal{S}$ is admissible, and then Lemma~\ref{lemma ID ==> (HA==>CDW)} shows that $\mathcal{S}$ is $\operatorname*{CDW}$-independent. \end{proof} Our final goal in this section is to prove that if $\left( \mathcal{C}% ,\mathcal{K}\right) $ is a connective island domain, then the systems of pre-islands are exactly the admissible subfamilies of $\mathcal{C}$. Recall that this is not true in general if $\left( \mathcal{C},\mathcal{K}\right) $ is not a connective island domain (see Example~\ref{example not CD}), but the two notions coincide for maximal families (Corollary~\ref{cor maximal HA <==> maximal SZ}). \begin{theorem} \label{thm ID ==> (HA<==>SZ)}If $\left( \mathcal{C},\mathcal{K}\right) $ is a connective island domain, then a subfamily of $\mathcal{C}$ is a system of pre-islands if and only if it is admissible. \end{theorem} \begin{proof} We have already seen in Proposition~\ref{prop SZ==>HA} that every system of pre-islands is admissible. Let us now assume that $\left( \mathcal{C}% ,\mathcal{K}\right) $ is a connective island domain and let $\mathcal{H}% \subseteq\mathcal{C}$ be admissible. From\ Lemma~\ref{lemma ID ==> (HA==>CDW)} it follows that $\mathcal{H}$ is $\operatorname*{CDW}$-independent. Let $\mathcal{S}$ be the system of pre-islands corresponding to $\left( \mathcal{C},\mathcal{K},h_{\mathcal{H}}\right) $, where $h_{\mathcal{H}}$ is the canonical height function of $\mathcal{H}$ (see Definition~\ref{def canonical height function}). Then $\mathcal{S}$ is also $\operatorname*{CDW}$-independent by Theorem~\ref{thm ID <==> (SZ==>CD)}. From Proposition~\ref{prop HA==>SZ resze} it follows that $\mathcal{H}% \subseteq\mathcal{S}$, and we are going to prove that we actually have $\mathcal{H}=\mathcal{S}$. Suppose for contradiction that there exists $S\in\mathcal{S}$ such that $S\notin\mathcal{H}$. Since $\mathcal{H}$ is $\operatorname{CD}$-independent and finite, the members of $\mathcal{H}$ that contain $S$ form a nonempty finite chain. Denoting the least element of this chain by $H$, we have $S\subset H$, as $S\notin\mathcal{H}$. Let $M_{1},\ldots,M_{m}$ denote those elements of $\mathcal{H}$ that are properly contained in $H$ and are maximal with respect to this property (if there are such sets). Clearly, $M_{1}% ,\ldots,M_{m}$ are pairwise disjoint, and $M_{1}\cup\cdots\cup M_{m}\subset H$, since $\mathcal{H}$ is $\operatorname*{CDW}$-independent (see Remark~\ref{rem CD maximal subsets}). We claim that $S\nsubseteq M_{1}\cup\cdots\cup M_{m}$. Assuming on the contrary that $S\subseteq M_{1}\cup\cdots\cup M_{m}$, the $\operatorname*{CDW}% $-independence of $\mathcal{S}$ implies that there is an $i\in\left\{ 1,\ldots,m\right\} $ such that $S\subseteq M_{i}$. However, this contradicts the minimality of $H$. Any two elements of $H\setminus\left( M_{1}\cup\cdots\cup M_{m}\right) $ are contained in exactly the same members of $\mathcal{H}$, therefore $h_{\mathcal{H}}$ is constant, say constant $c$, on this set (see Figure~\ref{fig proof}; cf. also Figure~\ref{fig kanonikus}). On the other hand, if $x\in M_{1}\cup\cdots\cup M_{m}$, then clearly we have $h_{\mathcal{H}}\left( x\right) \geq c$, hence $\min h_{\mathcal{H}}\left( H\right) =c$. Since $S$ is not covered by the sets $M_{i}$, it contains a point $u$ from $H\setminus\left( M_{1}\cup\cdots\cup M_{m}\right) $, therefore $\min h_{\mathcal{H}}\left( S\right) =h\left( u\right) =c$. Thus we have $S\subset H\in\mathcal{K}$ and $\min h_{\mathcal{H}}\left( S\right) =\min h_{\mathcal{H}}\left( H\right) $, contradicting that $S$ is a pre-island with respect to $\left( \mathcal{C},\mathcal{K},h_{\mathcal{H}% }\right) $. \end{proof} \begin{figure} [tb] \begin{center} \includegraphics[ natheight=5.446500in, natwidth=7.871500in, height=4.2514cm, width=6.1162cm ]% {proofb.pdf}% \caption{Illustration to the proof of Theorem~\ref{thm ID ==> (HA<==>SZ)}}% \label{fig proof}% \end{center} \end{figure} The maximum number of (pre-)islands certainly depends on the structure of the island domain $\left( \mathcal{C},\mathcal{K}\right) $. H\"{a}rtel \cite{Hartel} proved that the maximum number of rectangular islands on a $1\times n$ board is $n$, and Cz\'{e}dli \cite{czedli} generalized this result by showing that the maximum number of rectangular islands on an $n\times m$ board is $\left\lfloor \left( mn+m+n-1\right) /2\right\rfloor $. Although these are the only cases where the exact value is known, there are estimates in several other cases \cite{BHH,HNP,HHNS,L3,P}. In full generality, we have the following upper bound. \begin{theorem} \label{thm maximum<=|U|}If $\left( \mathcal{C},\mathcal{K}\right) $ is a connective island domain and $\mathcal{S}$ is a system of pre-islands corresponding to $\left( \mathcal{C},\mathcal{K}\right) $, then $\left\vert S\right\vert \leq\left\vert U\right\vert $. \end{theorem} \begin{proof} Let $\left( \mathcal{C},\mathcal{K}\right) $ be a connective island domain and let $\mathcal{S}\subseteq\mathcal{C}\setminus\left\{ \emptyset\right\} $ be a system of pre-islands corresponding to $\left( \mathcal{C}% ,\mathcal{K}\right) $. By Theorem~\ref{thm ID <==> (SZ==>CD)}, $\mathcal{S}$ is $\operatorname*{CDW}$-independent, and hence $\mathcal{S}\cup\left\{ \emptyset\right\} $ is also $\operatorname*{CDW}$-independent. From the results of \cite{CzSch} it follows that every maximal $\operatorname*{CDW}% $-independent subset of $\mathcal{P}\left( U\right) $ has $\left\vert U\right\vert +1$ elements. Thus we have $\left\vert \mathcal{S}\right\vert +1\leq\left\vert U\right\vert +1$. \end{proof} Observe that the above mentioned result of H\"{a}rtel shows that the bound obtained in Theorem~\ref{thm maximum<=|U|} is sharp. \section{Islands and proximity domains\label{section strict}} In this section we investigate islands, and we give a characterization of systems of islands corresponding to island domains $\left( \mathcal{C}% ,\mathcal{K}\right) $ satisfying certain natural conditions. We define a binary relation $\delta\subseteq\mathcal{C}\times\mathcal{C}$ that expresses the fact that a set $B\in\mathcal{C}$ is in some sense close to a set $A\in\mathcal{C}$:% \begin{equation} A\delta B\Leftrightarrow\exists K\in\mathcal{K}:~\text{ }A\preceq K\text{ and }K\cap B\neq\emptyset. \label{eq def delta}% \end{equation} \begin{remark} Let us note that the relation $\delta$ is not always symmetric. As an example, consider a directed graph, and let $\mathcal{C}=\mathcal{K}$ consist of $U$ and of those sets $S$ of vertices that have a source. (By a source of a set $S$ we mean a vertex $s\in S$ from which all other vertices of $S$ can be reached by a directed path that lies entirely in $S$.) It is easy to verify that in the graph $a\rightarrow b\rightarrow c\leftarrow d\leftarrow e$ we have $A\delta B$ but not $B\delta A$ for the sets $A=\left\{ a,b\right\} $ and $B=\left\{ c,d\right\} $. \end{remark} \begin{definition} \label{def distant}We say that $A,B\in\mathcal{C}$ are \emph{distant} if neither $A\delta B$ nor $B\delta A$ holds. Obviously, in this case $A$ and $B$ are also incomparable (in fact, disjoint), whenever $A,B\neq \emptyset$. A nonempty family $\mathcal{H}\subseteq\mathcal{C}$ will be called a \emph{distant family}, if any two incomparable members of $\mathcal{H}$ are distant. \end{definition} \begin{remark} \label{rem our delta is almost proximity}It is not difficult to verify that relation $\delta$ satisfies the following properties for all $A,B,C\in\mathcal{C}$ whenever $B\cup C \in \mathcal{C}: $ % \begin{align*} A\delta B & \Rightarrow B\neq\emptyset;\\ A\cap B\neq\emptyset & \Rightarrow A\delta B;\\ A\delta(B\cup C) & \Leftrightarrow(A\delta B\text{ or }A\delta C). \end{align*} \end{remark} \begin{lemma} \label{lemma distant ==> (CDW & HA)}If $\mathcal{H\subseteq C}$ is a distant family, then $\mathcal{H}$ is $\operatorname*{CDW}$-independent. Moreover, if $U\in\mathcal{H}$, then $\mathcal{H}$ is admissible. \end{lemma} \begin{proof} Let $\mathcal{H}\subseteq\mathcal{C}$ be a distant family, then $\mathcal{H}$ is clearly $\operatorname{CD}$-independent; moreover, it is easy to show using Remark~\ref{rem CD maximal subsets} that $\mathcal{H}$ is $\operatorname*{CDW}% $-independent. Next let us assume that $U\in\mathcal{H}$; we shall prove that $\mathcal{H}$ is admissible. Let $\mathcal{A}\subseteq\mathcal{H}$ be an antichain and let $H\in\mathcal{A}$. If $K\in\mathcal{K}$ contains $H$ properly, then there is a cover $K_{1}\in\mathcal{K}$ of $H$ such that $H\prec K_{1}\subseteq K$. Since all members of $\mathcal{A}\setminus\left\{ H\right\} $ are distant from $H$, none of them can intersect $K_{1}$, and therefore we have $K_{1}% \nsubseteq\bigcup\,\mathcal{A}$, and hence $K\nsubseteq\bigcup\,\mathcal{A}$. \end{proof} \begin{remark} \label{rem distant ==> standard h}Note that we have proved that $\mathcal{H}$ satisfies (\ref{eq stronger admissible}) for every antichain $\mathcal{A\subseteq H}$. Thus $h_{\mathcal{H}}$ is the standard height function of $\mathcal{H}$. \end{remark} \begin{theorem} \label{thm ID ==> (distant==>strSZ)}Let $\left( \mathcal{C},\mathcal{K}% \right) $ be a connective island domain and let $\mathcal{H}\subseteq \mathcal{C}\setminus\left\{ \emptyset\right\} $ with $U\in\mathcal{H}$. If $\mathcal{H}$ is a distant family, then $\mathcal{H}$ is a system of islands; moreover, $\mathcal{H}$ is the system of islands corresponding to its standard height function. \end{theorem} \begin{proof} Let $\mathcal{H}\subseteq\mathcal{C}\setminus\left\{ \emptyset\right\} $ be a distant family such that $U\in\mathcal{H}$. Applying Lemma~\ref{lemma distant ==> (CDW & HA)} we obtain that $\mathcal{H}$ is admissible, hence $\mathcal{H}$ is the system of pre-islands corresponding to $\left( \mathcal{C},\mathcal{K},h_{\mathcal{H}}\right) $ by Theorem~\ref{thm ID ==> (HA<==>SZ)}. Moreover, $h_{\mathcal{H}}$ is the standard height function of $\mathcal{H}$ by Remark~\ref{rem distant ==> standard h}. To finish the proof, we will prove that each $H\in\mathcal{H}$ is actually an island with respect to $\left( \mathcal{C},\mathcal{K},h_{\mathcal{H}% }\right) $. Suppose that $K\in\mathcal{K}$ is a cover of $H$. The distantness of $\mathcal{H}$ implies that the only members of $\mathcal{H}$ that intersect $K\setminus H$ are the ones that properly contain $H$. Since $h_{\mathcal{H}}$ is the standard height function, $h_{\mathcal{H}}\left( u\right) <\min h_{\mathcal{H}}\left( H\right) $ follows for all $u\in K\setminus H$. \end{proof} \begin{definition} \label{def proximity domain}The island domain $\left( \mathcal{C}% ,\mathcal{K}\right) $ is called a \emph{proximity domain}, if it is a connective island domain and the relation $\delta$ is symmetric for nonempty sets, that is \begin{equation} \forall A,B\in\mathcal{C}\setminus\left\{ \emptyset\right\} :~A\delta B\Leftrightarrow B\delta A. \label{eq delta symmetric}% \end{equation} \end{definition} If a relation $\delta$ defined on $\mathcal{P}\left( U\right) $ satisfies the three properties of Remark~\ref{rem our delta is almost proximity} and $\delta$ is symmetric for nonempty sets, then $\left( U,\delta\right) $ is called a \emph{proximity space}. The notion apparently goes back to Frigyes Riesz \cite{Riesz}, however this axiomatization is due to Vadim~A.~Efremovich (see \cite{E}). \begin{proposition} \label{prop PD ==> (strSZ==>distant)}If $\left( \mathcal{C},\mathcal{K}% \right) $ is a proximity domain, then any system of islands corresponding to $\left( \mathcal{C},\mathcal{K}\right) $ is a distant system. \end{proposition} \begin{proof} Let $\left( \mathcal{C},\mathcal{K}\right) $ be a proximity domain, and let $\mathcal{S}$ be the system of islands corresponding to $\left( \mathcal{C},\mathcal{K},h\right) $ for some height function $h$. Since $\left( \mathcal{C},\mathcal{K}\right) $ is a connective island domain, $\mathcal{S}$ is $\operatorname{CD}$-independent according to Theorem~\ref{thm ID <==> (SZ==>CD)}. Therefore, if $A,B\in\mathcal{S}$ are incomparable, then we have $A\cap B=\emptyset$. Assume for contradiction that $A\delta B$, i.e. that there is a set $K\in\mathcal{K}$ such that $A\prec K$ and $B\cap K\neq\emptyset$. Since $A$ and $B$ are disjoint, there exists an element $b\in\left( B\cap K\right) \setminus A$. Similarly, as we have $B\delta A$ by (\ref{eq delta symmetric}), there exists an element $a\in\left( A\cap K^{\prime}\right) \setminus B$ for some $K^{\prime}% \in\mathcal{K}$ with $B\prec K^{\prime}$. By making use of the fact that both $A$ and $B$ are islands with respect to $\left( \mathcal{C},\mathcal{K},h\right) $, we obtain the following contradicting inequalities:% \begin{align*} h\left( b\right) & <\min h\left( A\right) \leq h\left( a\right) ;\\ h\left( a\right) & <\min h\left( B\right) \leq h\left( b\right) .% \qedhere \end{align*} \end{proof} From Theorem~\ref{thm ID ==> (distant==>strSZ)} and Proposition~\ref{prop PD ==> (strSZ==>distant)} we obtain immediately the following characterization of systems of islands for proximity domains. \begin{corollary} \label{cor PD ==> (distant<==>strSZ)}If $\left( \mathcal{C},\mathcal{K}% \right) $ is a proximity domain, and $\mathcal{H}\subseteq\mathcal{C}% \setminus\left\{ \emptyset\right\} $ with $U\in\mathcal{H}$, then $\mathcal{H}$ is a system of islands if and only if $\mathcal{H}$ is a distant family. Moreover, in this case $\mathcal{H}$ is the system of islands corresponding to its standard height function. \end{corollary} Finally, let us consider the following condition on $\left( \mathcal{C}% ,\mathcal{K}\right) $, which is stronger than that of being a connective island domain:% \begin{equation} \forall K_{1},K_{2}\in\mathcal{K}:~K_{1}\cap K_{2}\neq\emptyset\implies K_{1}\cup K_{2}\in\mathcal{K}. \label{eq KK}% \end{equation} Observe that if we have a graph structure on $U$, and $\left( \mathcal{C}% ,\mathcal{K}\right) $ is a corresponding island domain (cf. Example~\ref{example graphs}), then (\ref{eq KK}) holds. \begin{theorem} \label{thm KK ==> proximity}Suppose that $\left( \mathcal{C},\mathcal{K}% \right) $ satisfies condition $\left( \ref{eq KK}\right) $, and assume that for all $C\in\mathcal{C}$, $K\in\mathcal{K}$ with $C\prec K$ we have $\left\vert K\setminus C\right\vert =1$. Then $\left( \mathcal{C}% ,\mathcal{K}\right) $ is a proximity domain, and pre-islands and islands corresponding to $\left( \mathcal{C},\mathcal{K}\right) $ coincide. Therefore, if $\mathcal{H}\subseteq\mathcal{C}\setminus\left\{ \emptyset \right\} $ and $U\in\mathcal{H}$, then $\mathcal{H}$ is a system of (pre-)islands if and only if $\mathcal{H}$ is a distant family. Moreover, in this case $\mathcal{H}$ is the system of (pre-)islands corresponding to its standard height function. \end{theorem} \begin{proof} Let $A,B\in\mathcal{C}\setminus\left\{ \emptyset\right\} $ such that $A\delta B$, i.e. $K\cap B\neq\emptyset$ for some $K\in\mathcal{K}$ with $A\preceq K$. If $A\cap B\neq\emptyset$, then clearly $B\delta A$ holds. Suppose now that $A\cap B=\emptyset$. By our assumption, $K=A\cup\left\{ b\right\} $ for some $b\in B$. From (\ref{eq KK}) it follows that $K\cup B\in\mathcal{K}$. Since $B\subset A\cup B=K\cup B\in\mathcal{K}$, there exists a cover $K^{\prime}\in\mathcal{K}$ of $B$ such that $B\prec K^{\prime }\subseteq A\cup B$. Clearly, we have $K^{\prime}\cap A\neq\emptyset$, hence $B\delta A$, and this proves that the relation $\delta$ is symmetric. Condition (\ref{eq KK}) is stronger than (\ref{eq def ID}), therefore $\left( \mathcal{C},\mathcal{K}\right) $ is a proximity domain. From our assumptions it is trivial that every pre-island with respect to $\left( \mathcal{C},\mathcal{K}\right) $ is also an island. The last two statements follow then from Corollary~\ref{cor PD ==> (distant<==>strSZ)}. \end{proof} \begin{corollary} Let $G$ be a graph with vertex set $U$; let $\left( \mathcal{C}% ,\mathcal{K}\right) $ be an island domain corresponding to $G$ (cf. Example~\ref{example graphs}), and let $\mathcal{H}\subseteq\mathcal{C}% \setminus\left\{ \emptyset\right\} $ with $U\in\mathcal{H}$. Then $\mathcal{H}$ is a system of (pre-)islands if and only if $\mathcal{H}$ is distant; moreover, in this case $\mathcal{H}$ is the system of (pre-)islands corresponding to its standard height function. \end{corollary} \section{Concluding remarks and an alternative framework} We introduced the notion of a (pre-)island corresponding to an island domain $\left( \mathcal{C},\mathcal{K}\right) $, where $U\in\mathcal{C}% \subseteq\mathcal{K}\subseteq\mathcal{P}\left( U\right) $ for a nonempty finite set $U$. We described island domains $\left( \mathcal{C}% ,\mathcal{K}\right) $ having $\operatorname{CD}$-independent systems of pre-islands, and we characterized systems of (pre-)islands for such island domains. In the general case, when no assumption is made on $\left( \mathcal{C},\mathcal{K}\right) $, we gave a necessary condition for a family of sets to be a system of pre-islands, and it remains an open problem to find an appropriate necessary and sufficient condition. Nevertheless, we obtained a complete characterization of \emph{maximal} systems of pre-islands in this general case. Determining the size of these maximal systems of pre-islands for specific island domains $\left( \mathcal{C},\mathcal{K}\right) $ has been, and continues to be, a topic of active research. Before concluding the paper, let us propose another possible approach to define islands. Let $U$ be a nonempty finite set and let $\mathcal{C}% \subseteq\mathcal{P}\left( U\right) $ with $U\in\mathcal{C}$, as before. We describe the \textquotedblleft surroundings\textquotedblright\ of members of $\mathcal{C}$ by means of a relation $\eta\subseteq U\times\mathcal{C}$, where $u\eta C$ means that the point $u\in U$ is close to the set $C\in\mathcal{C}$. We require $\eta$ to satisfy the following very natural axiom:% \begin{equation} \forall u\in U~\forall C\in\mathcal{C}:~u\in C\implies u\eta C. \label{eq eta axiom}% \end{equation} Examples of such \textquotedblleft point-to-set\textquotedblright\ proximity relations include closure systems (in particular, topological spaces) with $u\eta C$ if and only if $u$ belongs to the closure of $C$, and graphs with $u\eta C$ if and only if $u$ belongs to the neighborhood of $C$. We shall call a pair $\left( \mathcal{C},\eta\right) $ satisfying (\ref{eq eta axiom}) an \emph{island domain}. For any $C\in\mathcal{C}$, the set $\partial C:=\left\{ u\in U\colon u\eta C\text{ and }u\notin C\right\} $ is the set of points that surround $C$ (note that this is \emph{not} the usual notion of boundary for topological spaces). Therefore, we define islands corresponding to $\left( \mathcal{C}% ,\eta\right) $ as follows: If $h\colon U\rightarrow\mathbb{R}$ is a height function and $S\in\mathcal{C}$, then we say that $S$ is an \emph{island with respect to }$\left( \mathcal{C},\eta,h\right) $, if $h\left( u\right) <\min h\left( S\right) $ holds for all $u\in\partial S$. This definition is similar in spirit to the definition of an island corresponding to an island domain $\left( \mathcal{C},\mathcal{K}\right) $; in fact, it is a generalization of it. To see this, let us consider a pair $\left( \mathcal{C},\mathcal{K}\right) $, and let us define $\eta\subseteq U\times\mathcal{C}$ as follows:% \[ u\eta C\iff\exists K\in\mathcal{K}:C\preceq K\text{ and }u\in K. \] It is easy to verify that the islands corresponding to $\left( \mathcal{C}% ,\eta\right) $ are exactly the islands corresponding to $\left( \mathcal{C},\mathcal{K}\right) $. Let us now briefly sketch how to adapt the definitions of admissibility, connective island domain and distantness to this setting. We shall say that $\mathcal{H}\subseteq\mathcal{C}\setminus\left\{ \emptyset\right\} $ is \emph{admissible}, if $U\in\mathcal{H}$, and for every antichain $\mathcal{A}\subseteq\mathcal{H}$ we have% \[ \exists H\in\mathcal{A}\text{ such that }\forall u\in U:~u\in\partial H\implies u\notin\bigcup\,\mathcal{A}. \] We call the pair $\left( \mathcal{C},\eta\right) $ a \emph{connective island domain} if% \[ \forall A,B\in\mathcal{C}:~\left( A\cap B\neq\emptyset\text{ and }B\nsubseteq A\right) \implies\exists u\in B\setminus A:u\eta A. \] To define distantness, we extend $\eta$ to a \textquotedblleft set-to-set\textquotedblright\ proximity relation $\delta\subseteq \mathcal{C}\times\mathcal{C}$: for $A,B\in\mathcal{C}$, let $A\delta B$ if and only if there exists a point $u\in B$ with $u\eta A$. Using this relation $\delta$, we can define distant families just as in Definition~\ref{def distant}. Most of the results of this paper remain valid with these new definitions, and the proofs require only minor and quite straightforward modifications. The only exceptions are Lemma~\ref{lemma distant ==> (CDW & HA)}, where we need the extra assumption that $\left( \mathcal{C},\eta\right) $ is a connective island domain, and Theorem~\ref{thm KK ==> proximity}, which cannot be interpreted in this framework, as it refers to $\mathcal{K}$. The following theorem summarizes the main results. \begin{theorem} \label{thm eta}Let $U$ be a nonempty finite set, let $\mathcal{C}% \subseteq\mathcal{P}\left( U\right) $ with $U\in\mathcal{C}$, and let $\eta\subseteq U\times\mathcal{C}$ satisfy $\left( \ref{eq eta axiom}\right) $.% \renewcommand{\theenumi}{(\roman{enumi})} \renewcommand{\labelenumi}{\theenumi}% \begin{enumerate} \item A family $\mathcal{H}\subseteq\mathcal{C}\setminus\left\{ \emptyset\right\} $ is contained in a system of islands if and only if $\mathcal{H}$ is admissible. \item A family $\mathcal{H}\subseteq\mathcal{C}\setminus\left\{ \emptyset\right\} $ is a maximal system of islands if and only if $\mathcal{H}$ is a maximal admissible family. \item The pair $\left( \mathcal{C},\eta\right) $ is a connective island domain if and only if all systems of islands are $\operatorname{CD}% $-independent (equivalently, $\operatorname*{CDW}$-independent). \item If $\left( \mathcal{C},\eta\right) $ is a connective island domain, then a family $\mathcal{H}\subseteq\mathcal{C}\setminus\left\{ \emptyset \right\} $ is a system of islands if and only if $\mathcal{H}$ is admissible. \item If $\left( \mathcal{C},\eta\right) $ is a connective island domain and the corresponding relation $\delta$ is symmetric, then a family $\mathcal{H}% \subseteq\mathcal{C}\setminus\left\{ \emptyset\right\} $ is a system of islands if and only if $\mathcal{H}$ is distant and $U\in\mathcal{H}$. Moreover, in this case $\mathcal{H}$ is the system of islands corresponding to its standard height function. \end{enumerate} \end{theorem} \begin{corollary} \label{cor eta}Let $G=\left( U,E\right) $ be a connected simple graph, let $\mathcal{C}\subseteq\mathcal{P}\left( U\right) $ be a family of connected subsets with $U\in\mathcal{C}$, and let us define $\eta\subseteq U\times\mathcal{C}$ by% \[ u\eta C\iff u\in C\text{ or }\exists v\in C:~uv\in E. \] Then the following three conditions are equivalent for any $\mathcal{H}% \subseteq\mathcal{C}\setminus\left\{ \emptyset\right\} $ with $U\in \mathcal{H}$:% \renewcommand{\theenumi}{(\roman{enumi})} \renewcommand{\labelenumi}{\theenumi}% \begin{enumerate} \item $\mathcal{H}$ is a system of islands corresponding to $\left( \mathcal{C},\eta\right) $. \item $\mathcal{H}$ is an admissibly family. \item $\mathcal{H}$ is a distant family. \end{enumerate} \noindent If these conditions hold, then $\mathcal{H}$ is the system of islands corresponding to its standard height function. \end{corollary} \begin{proof} The fact that $\mathcal{C}$ contains only connected sets ensures that $\left( \mathcal{C},\eta\right) $ is a connective island domain, and it is trivial that $\delta$ is symmetric, hence we can apply Theorem~\ref{thm eta}. \end{proof} Let us note that in Corollary~\ref{cor eta} distantness of two sets $A,B\in\mathcal{C}$ means that there is no edge with one endpoint in $A$ and the other endpoint in $B$. Applying this corollary to a square grid (on a rectangular, cylindrical or toroidal board) or to a triangular grid, and letting $\mathcal{C}$ consist of all rectangles, squares or triangles, we obtain the earlier dry characterizations of islands as special cases. \subsection*{Acknowledgments} S\'{a}ndor Radeleczki acknowledges that this research was carried out as part of the TAMOP-4.2.1.B-10/2/KONV-2010-0001 project supported by the European Union, co-financed by the European Social Fund. Eszter K. Horv\'{a}th and Tam\'{a}s Waldhauser acknowledge the support of the Hungarian National Foundation for Scientific Research under grant no. K83219. Supported by the European Union and co-funded by the European Social Fund under the project ``Telemedicine-focused research activities on the field of Matematics, Informatics and Medical sciences'' of project number ``T\'AMOP-4.2.2.A-11/1/KONV-2012-0073'' Stephan Foldes acknowledges that this work has been co-funded by Marie Curie Actions and supported by the National Development Agency (NDA) of Hungary and the Hungarian Scientific Research Fund (OTKA, contract number 84593), within a project hosted by the University of Miskolc, Department of Analysis. The work was also completed as part of the TAMOP-4.2.1.B.- 10/2/KONV-2010-0001 project at the University of Miskolc, with support from the European Union, co-financed by the European Social Fund. \bigskip \bigskip \includegraphics[height=0.9cm]{logo_eu.pdf} \quad \includegraphics[height=1.2cm]{logo_marie-curie.pdf} \quad \includegraphics[height=1.2cm]{logo_nfu.pdf} \quad\includegraphics[height=1cm]{logo_otka.pdf} \bigskip
\section{Introduction} \label{intro} Whereas most neutron stars are endowed with typical magnetic fields of order $10^{12}$~G, a few of them have been found to have much stronger fields. Huge fields could be generated via dynamo effects in hot newly-born neutron stars with initial periods of a few milliseconds. Soft-gamma repeaters and anomalous X-ray pulsars are expected to be the best candidates of these so called \textit{magnetars}~\citep{woods2006}. In particular, a surface magnetic field of about $2.4\times 10^{15}$~G has been inferred in SGR~1806$-$20 and its internal field could be as high as $10^{18}$~G~\citep{lai91}. The presence of such strong fields makes the properties of magnetar crusts very different from those of ordinary neutron stars. \section{Structure and equation of state of magnetar crusts} \label{eos} We have studied the structure of the outer crust of a cold non-accreting magnetar using the model of ~\cite{lai91}. For this purpose, we have made use of the most recent experimental data from a preliminary unpublished version of an updated Atomic Mass Evaluation. For the atomic masses that have not yet been measured, we have employed the microscopic model HFB-21 of ~\cite{goriely2010}. In a strong magnetic field, the electron motion perpendicular to the field is quantized into Landau levels. As a result, the equilibrium composition of magnetar crusts can significantly differ from that of ordinary neutron stars, especially when the magnetic field strength $B$ exceeds $B_c\equiv m_e^2 c^3/(e\hbar)\simeq 4.4\times 10^{13}$~G~\citep{rila2011}. For instance, $^{66}$Ni which is found in the outer crust of neutron stars for $B=0$~\citep{pearson2011} disappear for $B>67 B_c$. On the other hand, $^{88}$Sr is only found in magnetar crusts for $B> 859 B_c$. Moreover, strong magnetic fields prevent neutrons from dripping out of nuclei. As a result, the pressure at the neutron drip transition increases from $7.82\times 10^{29}$ dyn~cm$^{-2}$ for $B=0$ to $1.05\times 10^{30}$ dyn~cm$^{-2}$ for $B=1000 B_c$. This might have implications for the interpretation of pulsar glitches~\citep{lrr}. As shown in Fig.~\ref{fig1}, the strongly quantizing magnetic fields prevailing in magnetar interiors have a large impact on the EoS in the regions where only a few Landau levels are filled. With increasing density, the effects of $B$ become less and less important as more and more levels are populated and the EoS matches smoothly with that obtained by~\cite{pearson2011} for $B=0$. \begin{figure} \centering \includegraphics[scale=0.33]{fig1} \caption{Pressure $P$ vs mass density $\rho$ in the outer crust of a cold non-accreting neutron star for different magnetic field strengths (in units of $B_c$). The filled squares indicate the points at which the lowest Landau level is fully occupied.} \label{fig1} \end{figure} \acknowledgments This work was supported by FNRS (Belgium), NSERC (Canada), Wallonie-Bruxelles-International (Belgium) and the Bulgarian Academy of Sciences.
\section{Introduction} The temporal quantum evolution of an isolated macroscopic system initially prepared in an out-of-equilibrium configuration is currently turning from an abstract concept, useful for discussing fundaments of quantum statistical mechanics, to a real phenomenon that can be observed and studied experimentally. This metamorphosis has been mainly driven by experiments on cold atoms,\cite{BlochRMP,SilvaRMP} but it will be surely given further impulse in the near future by the fast progresses in time-resolved spectroscopy on condensed-matter systems. In fact, the early time dynamics (up to $\sim 1$ ns) of a material that is driven out-of-equilibrium e.g. by an intense ultra-short laser pulse is still uninfluenced by the environmental heat sink, hence it is to a good approximation the dynamics of an isolated system. An interesting class of experiments focuses on the dynamics across phase transitions. In cold atom systems, this can be achieved by a sudden change of the experimental conditions, e.g. the depth of the optical lattice, which corresponds to suddenly altering the Hamiltonian parameters, a unique opportunity offered by these systems.\cite{Dynamics-1} In pump-probe experiments on real materials, one can instead tune the fluence of the pumping laser, i.e. the excess energy injected into the system.\cite{Nasu} That energy is supposed to undergo fast redistribution among all degrees of freedom, first among the electrons (faster), later among the phonons (slower), thus effectively heating the sample and raising its temperature. If the equilibrium phase diagram has a transition between a low temperature phase and a high temperature one, the effective temperature rise could drive such a phase transition, though the system will eventually relax back to its initial equilibrium state by the coupling to the external thermostat. In reality, a dynamical phase transition in an isolated macroscopic system is not as trivial as one could imagine. Across a thermodynamic transition, ergodicity is either lost or recovered, hence it is not at all obvious that the unitary time-evolution of an initial non-equilibrium quantum state should bring about the same results as, at equilibrium, the adiabatic change of a coupling constant or of temperature. This is actually a central question at the basis of quantum statistical mechanics. Here, more modestly, we highlight a possible link between the ergodicity breakdown at a phase transition and the occurrence in the many-body eigenvalue spectrum of "broken-symmetry edges", namely of special energies that mark the boundaries between symmetry-breaking and symmetry-invariant eigenstates. \begin{figure}[thb] \vspace{0.2cm} \centerline{\includegraphics[width=6cm]{Fig1.eps}} \caption{(Color online) Generic phase diagram of a model that possesses a symmetry broken phase in the temperature $T$ vs. coupling constant $g$ phase diagram, separated from a symmetric phase by a second order phase transition.} \label{Z2} \end{figure} Let us imagine a system described by a Hamiltonian $\mathcal{H}$, which undergoes a quantum phase transition, the zero-temperature endpoint of a whole second-order critical line that separates a low-temperature broken-symmetry phase from a high-temperature symmetric one, see Fig.~\ref{Z2} where $g$ is the coupling constant that drives the phase transition. For the sake of simplicity, and also because it will be relevant later, let us assume that the broken symmetry is a discrete $Z_2$ - all arguments below do not depend on this specific choice - with order parameter \begin{equation} \langle \sigma \rangle = \frac{1}{V}\sum_i\,\langle \sigma_i\rangle\in [-1,1],\label{OP} \end{equation} which is not a conserved quantity, i.e. $\Big[\sigma,\mathcal{H}\Big]\not =0$, and where $V\to\infty$ is the volume and $i$ labels lattice sites. Below the quantum critical point $g<g_c$, Fig. \ref{Z2}, the ground state is doubly degenerate and not $Z_2$ invariant. If $\mid \Psi_{\pm}\rangle$ are the two ground states, they can be chosen such that \[ \langle \Psi_{\pm}\mid\sigma\mid\Psi_{\pm}\rangle = \pm m,\qquad \text{with } m>0. \] On the contrary, for $g>g_c$, the ground state is unique and symmetric, i.e. the average of $\sigma$ vanishes. Since the symmetry breaking survives at finite temperature, see Fig. \ref{Z2}, one must conclude that, besides the ground state, a whole macroscopic set of low energy states is $Z_2$ not-invariant. The ergodicity breakdown in a symmetry broken phase specifically implies that these states, in the example we are dealing with, are grouped into two subspaces that are mutually orthogonal in the thermodynamic limit, one that can be chosen to include all eigenstates with $\langle \sigma\rangle >0$, the other those with $\langle \sigma\rangle <0$. Since the symmetry is recovered above a critical temperature, then there should exist a high energy subspace that includes symmetry invariant eigenstates. We argue that there should be a special energy in the spectrum, a {\sl "broken-symmetry edge"} $E_*$, such that all eigenstates with $E<E_*$ break the symmetry, while all eigenstates above $E_*$ are symmetric. In the case where the Hamiltonian has additional symmetries besides $Z_2$, hence conserved quantities apart from energy, we claim that, within each subspace invariant under these further symmetries, there must exist an edge above which symmetry is restored, even though its value may differ from one subspace to another as is the case in the model discussed in the next section. Let us for instance focus on any of these subspaces. The $Z_2$ symmetry implies that all eigenstates are even or odd under $Z_2$. The order parameter $\sigma$ is odd, hence its average value is strictly zero on any eigenstate, either odd or even. Nevertheless, we can formally define an order parameter $m(\Psi_E)$ of a given eigenstate $\mid\Psi_E\rangle$ through the positive square root of \begin{equation} m\big(\Psi_E\big) = \sqrt{\lim_{|i-j|\to\infty} \left|\langle \Psi_E\mid \sigma_i\sigma_j \mid \Psi_E\rangle\right|}.\label{OP-1} \end{equation} We denote by \begin{equation} \rho_{SB}(E) = \text{e}^{V\,S_{SB}(\epsilon)}, \end{equation} the density of symmetry-breaking eigenstates, namely those with $m(\Psi_E)>m_0$, where $m_0$ is a cut-off value that vanishes sufficiently fast as $V\to\infty$, being $\epsilon=E/V$ and $S_{SB}(\epsilon)$ their energy and entropy per unit volume. Seemingly, we define \begin{equation} \rho_{SI}(E)=\text{e}^{V\,S_{SI}(\epsilon)}, \end{equation} the density of the symmetry-invariant eigenstates, $m(\Psi_E)\leq m_0$, with $S_{SI}(\epsilon)$ their entropy. We claim that there exists an energy $E_*=V\epsilon_*$ that marks the microcanonical continuous phase transition in that specific invariant subspace, such that \begin{eqnarray} \lim_{V\to\infty} S_{SI}(\epsilon) &=& 0,\qquad \text{for }\epsilon<\epsilon_*, \label{S_SI}\\ \lim_{V\to\infty} S_{SB}(\epsilon) &=& 0,\qquad \text{for }\epsilon>\epsilon_*. \label{S_SB} \end{eqnarray} We do not have a rigorous proof of the above statement, but just a plausible argument. Let us suppose to define the average $m(E)>m_0$ of the order parameter over the symmetry-breaking eigenstates through \begin{equation} m(\epsilon) = \fract{1}{\rho_{SB}(E)}\sum_{\Psi_{E'}} \, m\big(\Psi_{E'}\big)\,\delta\big(E'-E\big).\label{m-SB} \end{equation} The actual microcanonical average is thus \begin{equation} \overline{m}(\epsilon) = \fract{\rho_{SB}(E)}{\rho_{SB}(E)+\rho_{SI}(E)}\, m(\epsilon).\label{m-ave} \end{equation} In the thermodynamic limit $V\to\infty$, hence $m_0\to 0$, the continuous phase transition would imply the existence of an energy $\epsilon_*$ such that, for $\epsilon\lesssim \epsilon_*$, $\overline{m}(\epsilon) \sim (\epsilon_*-\epsilon)^{\beta'}$, where the exponent $\beta'$ may not coincide with the corresponding one in the canonical ensemble $\overline{m}(T) \sim (T_c-T)^\beta$, while $\overline{m}(\epsilon>\epsilon_*)=0$. Since the entropy ratio on the r.h.s. of Eq. \eqn{m-ave} is either 1 or 0 in the thermodynamic limit, we conclude that the critical behavior comes from $m(\epsilon\lesssim\epsilon_*)\sim (\epsilon_*-\epsilon)^{\beta'}$, which, by continuity, implies $m(\epsilon>\epsilon_*)=0$, namely that there are no symmetry-breaking eigenstates with finite entropy density above $\epsilon_*$, hence Eq. \eqn{S_SB}. This further suggests that symmetry-breaking and symmetry-invariant eigenstates exchange their role across the transition, which makes also Eq. \eqn{S_SI} plausible. We do not exclude that symmetry-breaking eigenstates may survive above $\epsilon_*$, or vice versa for symmetric ones; we just state that, if they survive, their entropy is not extensive. We may also guess a generalization of the above picture to the most common situation of a first order phase transition. In this case we expect two different edges, $\epsilon_{1}<\epsilon_{*}$. Below $\epsilon_{1}$ the entropy density of symmetry-invariant states $S_{SI}(\epsilon)$ vanishes in the thermodynamic limit, while above $\epsilon_{*}$, the actual edge for symmetry restoration, it is $S_{SB}(\epsilon)$ that goes to zero. If we accept the existence of such an energy threshold, then we are also able to justify, without invoking any thermalization hypothesis,\cite{thermalization-1,thermalization-2} why a material, whose equilibrium phase diagram is like that of Fig. \ref{Z2}, may undergo a dynamical phase transition once supplied initially with enough excess energy so as to push it above $E_*$. We mention once more that the above arguments are not at all a real proof. However, they can be explicitly proven in mean-field like models, like the fully connected Ising model that we discuss in section \ref{Ising}. There, we explicitly demonstrate that the dynamical transition occurs because above a threshold energy there are simply no more broken-symmetry eigenstates in the spectrum. We believe this is important because it may happen that such an energy threshold, hence such a dynamical transition, exists also in models whose phase diagram is different from that of Fig. \ref{Z2}, as we are going to discuss in section \ref{Hubbard}. \section{First model: the fully connected Ising in a transverse field} \label{Ising} We consider the Hamiltonian of an Ising model in a transverse field \begin{eqnarray} \mathcal{H} &=& -\sum_{i,j}\,J_{ij}\,\sigma^z_i \sigma^z_j - h \sum_i\,\sigma^x_i\nonumber\\ &=& -\frac{1}{N}\sum_\mathbf{q}\, J_\mathbf{q}\,\sigma^z_\mathbf{q} \sigma^z_{-\mathbf{q}} - h\,\sigma^x_{\mathbf{0}},\label{Ham-0} \end{eqnarray} where $N$ is the number of sites, \[ \sigma^a_\mathbf{q} = \sum_i\,\text{e}^{-i\mathbf{q}\cdot\mathbf{r}_i}\,\sigma^a_i, \] is the Fourier transform of the spin operators, and $J_\mathbf{q}$ the Fourier transform of the exchange. In the (mean-field) fully-connected limit, $J_\mathbf{q} = J\,\delta_{\mathbf{q}\mathbf{0}}$, the model \eqn{Ham-0} simplifies into \begin{equation} \mathcal{H} = -\frac{1}{N}\,\sigma^z_\mathbf{0} \sigma^z_\mathbf{0} - h\,\sigma^x_\mathbf{0} = -\frac{4}{N} S^z S^z - 2h\,S^x,\label{Ham} \end{equation} having set $J=1$ and defined the total spin $\mathbf{S}=\boldsymbol{\sigma}_\mathbf{0}/2$. It turns out that the Hamiltonian Eq.~\eqn{Ham} can be solved exactly. We shall closely follow the work by Bapst and Semerjian,\cite{Semerjian} whose approach fits well our purposes. For reader's convenience we will repeat part of Bapst and Semerjian's calculations. We start by observing that the Hamiltonian \eqn{Ham} commutes with the total spin operator $\mathbf{S}\cdot\mathbf{S}$, with eigenvalue $S(S+1)$, so that one can diagonalize $\mathcal{H}$ within each $S\in [0,N/2]$ sector, which contains $2S+1$ distinct eigenvalues, each one $g(S)$ times degenerate, where $g(N/2)=1$ and, for $S<N/2$, \begin{equation} g(S)=\binom{N}{\fract{N}{2}+S} - \binom{N}{\fract{N}{2}+S+1}, \label{g(S)} \end{equation} which is the number of ways to couple $N$ spin-1/2 to obtain total spin $S$. We define \begin{equation} S = N\Big(\frac{1}{2} - k\Big) ,\label{p} \end{equation} where $k$, for large $N$, becomes a continuous variable $k\in [0,1/2]$. For a given $S$, a generic eigenfunction can be written as \begin{equation} \mid \Phi_E\rangle = \sum_{M=-S}^S\,\Phi_E(M)\mid M\rangle,\label{Psi} \end{equation} where $\mid M\rangle$ is eigenstate of $S^z$ with eigenvalue $M\in [-S,S]$. One readily find the eigenvalue equation\cite{Semerjian} \begin{eqnarray} E\,\Phi_E(M) &=& - \frac{4}{N}\,M^2\,\Phi_E(M) \label{eigen-1}\\ && - h\bigg[ \sqrt{S(S+1) - M(M-1)}\;\Phi_E(M-1) \nonumber\\ && ~~~+ \sqrt{S(S+1) - M(M+1)}\;\Phi_E(M+1)\bigg].\nonumber \end{eqnarray} We now assume $N$ large keeping $k$ constant. We also define \[ m = \frac{2M}{N} \in [-1+2k,1-2k], \] so that, at leading order in $N$, after setting $E=N\epsilon$ and \[ \Phi_E(M) = \Phi_\epsilon(m), \] the Eq.~\eqn{eigen-1} reads \begin{eqnarray} \epsilon\,\Phi_\epsilon(m) &=& - m^2\,\Phi_\epsilon(m) - \frac{h}{2}\,\sqrt{(1-2k)^2-m^2}\nonumber\\ && \;\Bigg[\Phi_\epsilon\left(m-\frac{2}{N}\right) + \Phi_\epsilon\left(m+\frac{2}{N}\right)\Bigg].\label{eigen-2} \end{eqnarray} Following Ref.\onlinecite{Semerjian}, we set \begin{equation} \Phi_\epsilon(m) \propto \exp\big[ -N\,\phi_\epsilon(m)\big],\label{ansatz} \end{equation} where the proportionality constant is the normalization, so that \begin{eqnarray*} \Phi_\epsilon\left(m\pm \frac{2}{N}\right) &\propto& \exp\bigg[ -N\,\phi_\epsilon\left(m\pm \frac{2}{N}\right)\bigg]\\ && \simeq \Phi_\epsilon(m)\,\text{e}^{\mp 2\,\phi'_\epsilon(m)}, \end{eqnarray*} Upon substituting the above expression into \eqn{eigen-2}, the following equation follows \begin{equation} \phi'_\epsilon(m) = \frac{1}{2}\,\text{arg}\,\cosh\bigg(- \fract{\epsilon+m^2}{h\sqrt{(1-2k)^2-m^2}}\bigg).\label{eigen-3} \end{equation} For large $N$, in order for the wave function \eqn{ansatz} to be normalizable, we must impose that: ({\sl i}) the $\Re \text{e} \,\phi_\epsilon(m) \geq 0$; ({\sl ii}) the $\Re \text{e} \,\phi_\epsilon(m)$ must have zeros, which, because of ({\sl i}), are also minima. As showed in Ref. \onlinecite{Semerjian}, these two conditions imply that the allowed values of the energy are \begin{equation} \text{min}\Big(f_-(m)\Big)\leq \epsilon \leq \text{Max}\Big(f_+(m)\Big),\label{c-1}, \end{equation} where \begin{eqnarray} f_+(m) &=& - m^2 + h\,\sqrt{(1-2k)^2-m^2},\label{f+}\\ f_-(m) &=& - m^2 - h\,\sqrt{(1-2k)^2-m^2} .\label{f-} \end{eqnarray} At fixed $k$, the lowest allowed energy is thus \begin{equation} \epsilon_\text{min}(k)=\text{min}\left(f_-(m)\right) = - (1-2k)^2 - \frac{h^2}{4},\label{e<-eq} \end{equation} and occurs at \begin{equation} m^2(k) = (1-2k)^2 - \frac{h^2}{4},\label{m-eq} \end{equation} if $h\leq h(k) = 2(1-2k)$, otherwise the minimum energy occurs at $m=0$, \begin{equation} \epsilon_\text{min}(k) = f_-(0) = -h\,(1-2k).\label{e>-eq} \end{equation} It follows that the actual ground state is always in the $k=0$ subspace and has energy \begin{equation} \epsilon_0 = \begin{cases} -1 - \frac{h^2}{4} & \mbox{if } h\leq h(0)=2,\\ -h & \mbox{if } h>h(0). \end{cases} \end{equation} In Fig. \ref{1} we plot the two functions $f_+(m)$ and $f_-(m)$ for $k=0$ and $h=0.9<h(0)$. \begin{figure}[t] \centerline{\includegraphics[width=7cm]{Fig2.eps}} \vspace{-0.8cm} \caption{\label{1} The two function $f_+(m)$ and $f_-(m)$ for $k=0$ and $h=0.9$. The allowed values of the eigenvalues are those between the minimum of $f_-$ and the maximum of $f_+$.} \end{figure} As shown by Bapst and Semerjian,\cite{Semerjian} whenever $f_-(m)$ has a double minimum as in Fig.~\ref{1}, any eigenstate with energy below $\epsilon < \epsilon_*=f_-(m=0)$ is doubly degenerate in the thermodynamic limit $N\to\infty$, being localized either at positive or at negative $m$, thus not invariant under $Z_2$. On the contrary, the eigenvalues for $\epsilon\geq \epsilon_*$ are not degenerate and are $Z_2$ symmetric. More specifically, any eigenfunction $\Phi_\epsilon(m)$ has evanescent tails that vanish exponentially with $N$ in the regions where $f_-(m) > \epsilon$ and $f_+(m) < \epsilon$. In Fig. \ref{evan} we show in the case $\epsilon<\epsilon_*$ the regions of evanescent waves. In this case, one can construct two eigenfunctions, each localized in a well, whose mutual overlap vanishes exponentially for $N\to\infty$. This result also implies that the ground state, which lies in the $k=0$ subspace, spontaneously breaks $Z_2$ when $h<h(0)$, hence $h(0)=2=h_c$ is the critical transverse field at which the quantum phase transition takes place. Such a degenerate ground state is actually a wave packet centered either at $m=+\sqrt{1-h^2/4}$ or at $-\sqrt{1-h^2/4}$, see Eq.~\eqn{m-eq}. \begin{figure}[bht] \centerline{\includegraphics[width=7cm]{Fig3.eps}} \vspace{-0.8cm} \caption{\label{evan} (Color online) For a given energy $\epsilon<\epsilon_*$, we draw in red the regions $\epsilon<f_-(m)$ where the wave function vanishes exponentially as $N\to\infty$.} \end{figure} More generally, it follows that for any given $k$ and $h<h(k)=2(1-2k)$ there is indeed a "broken-symmetry edge" \begin{equation} \epsilon_*(k) = -h\,(1-2k),\label{mobility-edge} \end{equation} that separates symmetry breaking eigenstates at $\epsilon<\epsilon_*$ from symmetric eigenstates at higher energies. In particular, in the lowest energy subspace with $k=0$, the edge is $\epsilon_*(0)=-h$. Therefore, although in the simple mean-field like model Eq.~\eqn{Ham}, one can indeed prove the existence of energy edges that separate symmetry invariant from symmetry breaking eigenstates. We also note that subspaces corresponding to different $k$ have different $\epsilon_*(k)$, as we anticipated in the Introduction. \subsection{The role of the broken-symmetry edge in the quench dynamics} \begin{figure}[ht] \centerline{\includegraphics[width=8cm]{Fig4.eps}} \vspace{-0.6cm} \caption{\label{quench} (Color online) Pictorial view of the quench dynamics. The upper curve (red) corresponds to the Hamiltonian for $t<0$ characterized by $h_i=0.8$. We assume that for negative time the wave function is the ground state for $m>0$, which corresponds to the minimum of the curve at positive $m_i$, dashed vertical line. The intermediate curve (blue) corresponds to $f_-(m)$ for $h_f=1.2$, while the lowest curve (black) to $h_f=1.5$. The intercepts $\epsilon=f_-(m_i)$ define the lower bound of allowed energies.} \end{figure} The quench dynamics we examine corresponds to propagating the ground state at $h=h_i$ with a different transverse field $h=h_f>h_i$. In the specific case of a fully-connected model, this problem has been addressed by Refs. \onlinecite{Sengupta} and \onlinecite{SciollaBiroli_long}. In particular, it has been found\cite{SciollaBiroli_long} that for $h_i<h_c$, i.e. starting from the broken-symmetry phase, a dynamical transition occurs at $h_f=h_* = (h_c+h_i)/2$. For $h_f\geq h_*$, the symmetry is dynamically restored, while, below, it remains broken as in the initial state. If, instead of a sudden increase from $h_i<h_c$ to $h_f$, one considers a linear ramp \[ h(t) = \begin{cases} h_i + (h_f-h_i)\,\fract{t}{\tau} & \text{for } t\in[0,\tau],\\ h_f & \text{for } t>\tau, \end{cases} \] then the critical $h_*$ increases and tends asymptotically to the equilibrium critical value $h_c(0)$ for $\tau\to\infty$,\cite{Sandri} as expected for an adiabatically slow switching rate. This result demonstrates that the dynamical transition is very much the same as the equilibrium one, and it occurs for lower fields $h$ only because of anti-adiabatic effects, which is physically plausible. The dynamical transition can be easily discovered in the semiclassical limit. If we set \begin{eqnarray*} \langle S^z\rangle &=& \frac{N}{2}\,\cos\theta,\\ \langle S^x\rangle &=& \frac{N}{2}\,\sin\theta\,\cos\phi,\\ \langle S^y\rangle &=& \frac{N}{2}\,\sin\theta\,\sin\phi, \end{eqnarray*} for large $N$ it is safe to assume $\langle S^z S^z\rangle \sim \langle S^z\rangle\langle S^z\rangle = N^2\cos^2\theta/4$, so that the total energy per spin, see Eq. \eqn{Ham}, reads \begin{equation} e = \fract{\langle \mathcal{H}\rangle}{N} = -\cos^2\theta - h\,\sin\theta\,\cos\phi,\label{e-semi} \end{equation} and is conserved in the unitary evolution. Through the Heisenberg equation $i\dot{S}^z = [S^z,\mathcal{H}]$ and upon expressing $\cos\phi$ as function of $e$ and $\theta$ by means of Eq. \eqn{e-semi}, one finds the equation of motion of the order parameter $m=\cos\theta$ \begin{equation} \fract{\partial \cos\theta}{\partial t} = \mp 2\sqrt{\left(h\sin\theta-e-\cos^2\theta\right)\left(h\sin\theta+e+\cos^2\theta\right)}. \end{equation} It follows that, until $e<-h$, the order parameter oscillates around a finite value -- the symmetry remains broken -- while for $e>-h$ it oscillates around zero -- the symmetry is restored. The dynamical transition thus occurs when $e_*=-h$, which we recognize to be the edge between symmetry-breaking and symmetry-invariant eigenstates at $k=0$ previously defined. This correspondence can be shown to hold also in the exact solution of the previous section. To this end, imagine to start from the ground state at an initial $h_i<h_c$, which occurs in the subspace of $k=0$, i.e. maximum total spin $S=N/2$, and let it evolve with the Hamiltonian at a different $h_f>h_i$. We note that, since $S$ is conserved, the time-evolved wave function will stay in the subspace $k=0$. The ground state at $h_i$ is degenerate, and we choose the state with a positive average of $m$, which is a wave-packet narrowly centered around $m_i = \sqrt{1-(h_i/2)^2}$. For very large $N$, when the contributions from the evanescent waves in the regions where $\Re \text{e} \phi_\epsilon(m) >0$ can be safely neglected, the initial wave function decomposes in $k=0$ eigenstates of the final Hamiltonian with eigenvalues $\epsilon$ such that $f_-(m_i)\leq \epsilon \leq f_+(m_i)$. In Fig. \ref{quench} we show graphically the condition $\epsilon\geq f_-(m_i)$ for $h_i=0.8$ and two values of $h_f=1.2, 1.5$. We note that the minimum value of the allowed energy for $h_f=1.2$ belongs to the subspace of symmetry broken states, while for $h_f=1.5$ it belongs to the subspace of symmetric states. It follows that, while for $h_f=1.2$ the long time average of $m$ will stay finite, for $h_f=1.5$ it will vanish instead. The critical $h_f=h_*$ is such that $f_-(m_i) = f_-(0) = -h_*$, namely $h_*=1+h_i/2 = (h_c+h_i)/2$, which is indeed the result of Ref. \onlinecite{SciollaBiroli_long}. Therefore, the dynamical restoration of the symmetry is intimately connected to the existence of an energy threshold. When the initial wave function decomposes into eigenstates of the final Hamiltonian that all have energies higher than that threshold, then the long time average of the order parameter vanishes although being initially finite. We note that the dynamical transition in this particular example is related to the equilibrium quantum phase transition, but it is actually unrelated to the transition at finite temperature.\cite{Semerjian} In fact, at a given value of the transverse field $h<h_c$, all eigenstates within the subspaces with $k\geq h_c -h$ are symmetric, while those with smaller $k$ have still low-energy symmetry-breaking eigenstates. Since the degeneracy $g(S)$, see Eq. \eqn{g(S)}, increases exponentially in $N$ upon lowering $S$, hence raising $k$, the entropic contribution of the symmetric subspaces at large $k$ will dominate the free energy and eventually drive the finite temperature phase transition. On the contrary, the quench-dynamics is constrained within the subspace at $k=0$, hence it remains unaware that in other subspaces the eigenstates at the same energy are symmetric. This observation is important and makes one wonders how the above result can survive beyond the fully-connected limit. Indeed, as soon as the Fourier transform of the exchange $J_\mathbf{q}$, see Eq. \eqn{Ham-0}, acquires finite components at $\mathbf{q}\not=\mathbf{0}$, states with different total spin, hence different $k$, start to be coupled one to each other -- the total spin ceases to be a good quantum number, the only remaining one being the total momentum. Therefore, symmetry-breaking eigenstates at low $k$ get coupled to symmetric eigenstates at large $k$. In this more general situation, there are to our knowledge no rigorous results apart from the pathological case of one-dimension, where the energy above which symmetry is restored actually coincides with the ground state energy, or, more rigorously, where excited states that breaks the symmetry do not have extensive entropy. However, we mention that a recent attempt to include small $\mathbf{q}\not=\mathbf{0}$ fluctuations on top of the results above, i.e. treating \begin{equation} -\frac{1}{N}\sum_{\mathbf{q}\not=\mathbf{0}}\, J_\mathbf{q}\,\sigma^z_\mathbf{q} \sigma^z_{-\mathbf{q}} \label{perturb} \end{equation} as a small perturbation of the {\sl bare} Hamiltonian \[ \mathcal{H}_0 = -\frac{J_\mathbf{0}}{N}\,\sigma^z_\mathbf{0} \sigma^z_{\mathbf{0}} - h\,\sigma^x_{\mathbf{0}}, \] suggests that the dynamical transition in the quench does survive,\cite{Sandri} which we take as an indirect evidence that the energy edge does, too. We suspect the reason being that, once symmetry-breaking and symmetry-invariant eigenstates of same energy get coupled by \eqn{perturb}, the new eigenstates, being linear combinations of the former ones, will all be symmetry-breaking. As a result, the broken-symmetry edge will end to coincide approximately with $\epsilon_*(k=0)$, i.e. with the maximum value among all the formerly independent subspaces. \section{A different example: the Hubbard model} \label{Hubbard} In the Introduction we inferred the existence of a threshold energy by the existence of a finite temperature phase transition that ends at $T=0$ into a quantum critical point. However, we found in the previous section an energy edge above which symmetry is restored that is actually unrelated to the finite temperature phase transition, while it is only linked to the quantum phase transition. This suggests that such an edge could exists in a broader class of situations. In some simple cases, one can actually prove its existence without much effort. Let us consider for instance the fermionic Hubbard model in three dimensions, with Hamiltonian \begin{equation} \mathcal{H} = -\sum_{i,j,\sigma}\, t_{ij}\,\Big(c^\dagger_{i\sigma}c^{\phantom{\dagger}}_{j\sigma} + H.c.\Big) + U\sum_i\, n_{i\uparrow}n_{i\downarrow},\label{Hubb} \end{equation} where $U>0$ is the on-site repulsion. In the parameter space where magnetism can be discarded, the low energy part of the spectrum is that of a normal metal, hence all eigenstates are expected to be invariant under the symmetries of the Hamiltonian $\mathcal{H}$, namely translations, spin-rotations and gauge transformations. We observe that the high-energy sector of the many-body spectrum obviously corresponds to the low energy spectrum of the Hamiltonian $-\mathcal{H}$. The latter is characterized by an opposite band dispersion but, more importantly, by an attractive rather than repulsive interaction. As a result of the Cooper instability, the low-energy spectrum of $-\mathcal{H}$ must comprise eigenstates $\mid \Psi\rangle$ with superconducting off-diagonal long range order, \begin{equation} \lim_{|i-j|\to\infty} \langle \Psi\mid d^\dagger_{i\uparrow}d^\dagger_{i\downarrow}\, d^{\phantom{\dagger}}_{j\downarrow}d^{\phantom{\dagger}}_{j\uparrow}\mid\Psi\rangle = \mid\Delta\mid^2>0, \label{ODLRO} \end{equation} which are not invariant under gauge transformations. Since these are identically the high-energy eigenstates of the original repulsive Hamiltonian Eq. \eqn{Hubb}, it follows that the upper part of its many-body spectrum contains eigenstates with off-diagonal long range order. There must thus exist a special energy which separates low-energy gauge-symmetric eigenstates from high-energy superconducting ones. Suppose to prepare a wave function that initially has $\mid\Delta(t=0)\mid^2>0$, see Eq. \eqn{ODLRO}, and let it evolve with the Hamiltonian $\mathcal{H}$, Eq. \eqn{Hubb}. If its energy is high enough, so that its overlap with the upper part of the spectrum is finite, then $\mid\Delta(t\to\infty)\mid^2>0$, otherwise $\mid\Delta(t\to\infty)\mid^2\to 0$, signaling once again a dynamical transition that, unlike the previous example of section \ref{Ising}, is now accompanied by the emergence rather than disappearance of long-range order. This surprising result was already conjectured by Rosch {\sl et al.} in Ref. \onlinecite{Rosch} as a possible metastable state attained by initially preparing a high energy wave function with all sites either doubly occupied or empty at very large $U$. What we have shown here is that such a superfluid behavior is robust and it is merely a consequence of the high energy spectrum of $\mathcal{H}$, which contains genuinely superconducting eigenstates. We finally observe that these states have negative temperature, hence are invisible in thermodynamics unless one could effectively invert the thermal population.\cite{Werner-U} \begin{figure}[hbt] \vspace{0.6cm} \centerline{\includegraphics[width=7.cm]{Fig5.eps}} \caption{Time evolution of the superconducting order parameter $\Delta(t)$ for a repulsive $U=0.2 U_c$, where $U_c$ is the critical repulsion at the Mott transition, and four different initial values $\Delta_0$, the curves a, b, c and d. We also show in the inset the early time relaxation of $\Delta(t)$ in case a. Time is is units of the inverse of half the bandwidth. We observe that the curves b, c and d maintain a finite order parameter, unlike the curve a, although it corresponds to the largest initial $\Delta_0=0.7$. We also note that in the case d with the lowest $\Delta_0=0.1$, the order parameter actually grows in time. } \label{Delta} \end{figure} While the existence of a high-energy subspace of superfluid eigenstates of the Hamiltonian \eqn{Hubb} is evident by the above discussion, it is worth showing explicitly its consequences in the out-of-equilibrium dynamics. To this end, we prepare an initial wave function $\mid \Psi(t=0)\rangle =\mid \Psi_\lambda\rangle$, ground state of the BCS Hamiltonian \begin{equation} \mathcal{H}_\text{BCS} = +\sum_{i,j,\sigma}\, t_{ij}\,\Big(c^\dagger_{i\sigma}c^{\phantom{\dagger}}_{j\sigma} + H.c.\Big) + \sum_i\, \Big(\lambda\,c^\dagger_{i\uparrow}c^\dagger_{i\downarrow} + H.c.\Big),\label{BCS} \end{equation} where the sign of the hopping is changed with respect to \eqn{Hubb}, and $\lambda$ is a control parameter that allows to span the alleged high-energy superconducting subspace of the repulsive Hubbard model by tuning the value of the initial order parameter \begin{equation} \Delta_0 \equiv \Delta(t=0) = \langle \Psi_{\lambda} \mid c^\dagger_{i\uparrow}c^\dagger_{i\downarrow} + c_{i\downarrow} c_{i\uparrow} \mid \Psi_{\lambda}\rangle. \label{order-parameter} \end{equation} Since this model proved insoluble so far, we resort to an approximate method for simulating the unitary time-evolution: the time-dependent Gutzwiller approximation.\cite{SchiroFabrizio_short,SchiroFabrizio_long} In brief, the time-evolved wave function is approximated by the form \begin{equation} \mid\Psi(t)\rangle \simeq \prod_i\,\mathcal{P}_i(t)\,\mid\Psi_{\lambda(t)}\rangle, \end{equation} where $\mathcal{P}_i(t)$ is a non-hermitian time-dependent variational operator that acts on the Hilbert space at site $i$, and $\mid\Psi_{\lambda(t)}\rangle$ the solution of the Schr{\oe}dinger equation with a BCS Hamiltonian like in Eq. \eqn{BCS} but with time-dependent coupling constants $t_{ij}$ and $\lambda$. \begin{figure}[hbt] \vspace{-0.4cm} \centerline{\includegraphics[width=7cm,angle=-90]{Fig6.eps}} \vspace{-.5cm} \caption{Long-time average of the order parameter, $\bar{\Delta}$, as function of its initial value, $\Delta_0$. In the inset we show the energy, in units of half-bandwidth, with respect to the ground state one obtained within the Gutzwiller approximation, as function of $\Delta_0$, as well as the broken-symmetry edge, the dashed line, corresponding to $U=0.2 U_c$. We note that the energy, i.e. the average value of the Hamiltonian \eqn{Hubb} on the ground state of \eqn{BCS}, is lower the greater $\Delta_0$ because of the sign change of the hopping in the two Hamiltonians.} \label{Delta-av} \end{figure} All variational parameters are determined through the saddle point of the action \begin{equation} \mathcal{S} = \int dt\, \langle\Psi(t)\mid i\frac{\partial}{\partial t} - \mathcal{H}\mid\Psi(t)\rangle, \label{action} \end{equation} which is computed within the Gutzwiller approximation. The method has been described in detail elsewhere, see for instance Ref. \onlinecite{Hvar}, and we just present the results here. For convenience, all calculations are done at half-filling, where they are much simpler, focusing on the paramagnetic sector, i.e. discarding magnetism, and assuming a flat density of states with half-bandwidth $D$ that we take as the energy unit. However, in order to avoid spurious interference effects from the Mott localization that occurs above a critical $U_c$, we concentrate on values of $U$ safely below $U_c$, in fact below the dynamical analogue of the Mott transition that is found when $U\gtrsim U_c/2$. \cite{WernerPRL,SchiroFabrizio_short} In fig. \ref{Delta} we plot the time evolution of the order parameter for $U=0.2 U_c$ and different initial values of $\Delta_0$, see Eq. \eqn{order-parameter}, as well as different energies, see inset of Fig. \ref{Delta-av}, where we also show the time-averaged values of $\Delta(t)$ with respect to their initial values. We observe that at the lower energies, which actually correspond to the larger $\Delta_0$, the order parameter, initially finite, relaxes rapidly to zero. On the contrary, above a threshold energy, $\Delta(t)$ stays finite and even grows with respect to its initial value, see Fig. \ref{Delta-av}. If we identify that threshold energy as our broken-symmetry edge, although rigorously it is only a lower bound, its dependence upon $U$ is shown in Fig. \ref{mobility-edge}. \begin{figure}[hbt] \vspace{0.6cm} \centerline{\includegraphics[width=8cm]{Fig7.eps}} \caption{Broken-symmetry edge measured with respect to the ground state energy for different values of $U$} \label{mobility-edge} \end{figure} The fermionic Hubbard model that we have discussed so far it is only a very simple example where one can infer the existence of a threshold energy in the many-body spectrum that is not accompanied by any anomaly in the thermodynamics. In fact, the idea of comparing the low-energy spectrum of a Hamiltonian $\mathcal{H}$ with that of $-\mathcal{H}$, which is actually the high-energy spectrum of $\mathcal{H}$, is very simple yet very effective. However, we must remark that model Hamiltonians $\mathcal{H}$, like the Hubbard model above, are meant to describe low energy properties of complex physical systems. Therefore, it is not unlikely that the value of the threshold energies extracted by comparing the low energy spectra of $\mathcal{H}$ and $-\mathcal{H}$ could be above the limit of applicability of the model itself, in which case they would be devoid of physical relevance. \section{Conclusions} The original purpose of this work was to understand how an isolated macroscopic quantum system brought away from equilibrium by a sudden injection of energy could cross a order-disorder phase transition as if its temperature were raised. We argued that this is possible because the eigenvalue spectrum of the system is characterized by a well defined energy, which we named {\sl broken-symmetry edge}, that separates low energy symmetry-breaking eigenstates from high-energy symmetry-invariant ones. Once the initial energy exceeds such a threshold, a dynamical restoration of symmetry occurs that mimics the equilibrium phase transition as obtained by increasing temperature or adiabatically changing coupling constants. We explicitly demonstrated this mechanism in the fully-connected Ising model in a transverse field, a prototypical mean-field example of a quantum phase transition. After that, we realized that such edges might arise also when the equilibrium phase diagram does not display any order-disorder phase transition, and bring about unexpected dynamical features. For instance, the repulsive Hubbard model has a high energy sector of genuinely superfluid eigenstates. By simulating in the Hubbard model the unitary evolution of a BCS wave function we surprisingly found that, if the energy exceeds a threshold, the initial superfluid order parameter not only survived in the long time limit, but may actually grew, despite the fact that the interaction is repulsive. In summary we believe that the concept of broken-symmetry edges that separate well distinct sectors of the eigenvalue spectrum is quite general and may be a useful framework of reference in the out-of-equilibrium quantum dynamics of isolated systems. \section*{Acknowledgments} We thank Erio Tosatti, Claudio Castellani and Marco Schir\`o for discussions and comments. This work has been supported by the European Union, Seventh Framework Programme, under the project GO FAST, grant agreement no. 280555.
\section{Introduction} The formation of a surface seal in soil, the hardening of the seal to form a crust and cracking of this crust, has been observed to occur under the combined influence of rain and dry weather \cite{1a, 2a}. The sealing leads to a local increase in the bulk density and a decrease in porosity along with a decrease in the hydraulic conductivity \cite{4a}. After drying these surface seals transform into surface crusts. The sealing and crusting of soil surfaces increases the runoff in the soil surfaces. In fact, initially the soil loss reduces because soil strength increases during crusting. However, crusting eventually leads to the formation of cracks which increases runoff in the soil surfaces \cite{5a}. The detailed effect of crusting on soil loss depends on various factors like the distribution of crusted and non-crusted regions and the shrinkage cracks within the crusted layer \cite{6a}. Cracking depends critically on the rate of drying and so it has been observed to occur within a few days of a rainstorm in a dry season \cite{6b}. The effects of sealing and crusting on soil loss are most visible in agricultural lands. This is because the sealing and crusting is most likely to occur in the soil which is not vegetated in the agricultural environment. In Europe a quarter of its agricultural land exhibits some form of soil erosion risk \cite{7ab}. In fact, a 20 mm rain has been observed to form surface seal and cause soil loss in certain hilly areas \cite{7a,7b}. In this case, the thickness of the surface seal was observed to be on the order of a few millimeters. However, the formation of a 7 cm thick surface seal during a wet season has also been observed \cite{8a}. Since the soil loss can effect the properties of soil and thus have a major impact on agricultural production, it is important to understand the physical processes at work in drying ground. In this paper we analyse convective instabilities that are observed to occur during the evaporation of a mixture of soil particles and water \cite{ee}. Evaporation of pure liquids can cause convection in those liquids \cite{a,b,c,d,e,f}. This is because the evaporation causes the surface temperature to drop and thus sets up an unstable density gradient which can induce a Rayleigh–Bernard instability. In addition, for most liquids surface tension is a function of temperature and so perturbations in temperature along the film surface also create perturbations in the surface tension. The liquid then flows from places of lower surface tension to places of higher surface tension. This flow is called the Marangoni effect. Thus, both the temperature gradient and surface tension variations contribute to the occurrence of convection in pure liquids. The combined phenomenon is called Bernard–Marangoni convection. The formation of surface seals and crusts can be understood by analysing the drying of pools of muddy water. Evaporation in mixtures also leads to convection \cite{aa,ba,ca,da}. In this case as the liquid drys up a concentration gradient is also set up along with a temperature gradient. The liquid moves under the combined influence of both these gradients. Furthermore, the Marangoni effect also occurs due to the dependence of the of surface tension on both the temperature and the concentration. Thus, a perturbation in either the temperature or the concentration can cause convection. Convection has been observed to occur in the course of evaporation of a binary mixture of soil particles and water \cite{ee}. In doing so a four weight percent mixture of bentonite and water was analysed at twenty degrees centigrade. It was found that convection occurred during the course of drying of this mixture and produced hexagonal patterns. However, it was also observed that the temperature gradient set up during the evaporation could not explain the occurrence of convection in this system. In this paper we will show that the concentration gradient is the main driving force behind such convection during the drying of ground. We also study the thawing of frozen ground which occurs in the Arctic circle. Freezing and thawing of soil in the Arctic circle results in the formation of various surface patterns such as soil hummocks and stone circles \cite{ar1, ar2, br1, br2}. Various models have been proposed to explain the occurrence of these patterns. In one of these models the maximum density of water at four degree centigrade sets up the convection \cite{ar4, ar41, ar5}. This model is based on the convection of water through soil pores which has not been observed for the soils under consideration. It has also been proposed that the sedimentation of soil during thawing sets up an unstable density profile which causes convection \cite{ara1, ara2, ara0}. However, the measured soil density gradient is not large enough to initiate convection \cite{ara4}. Another model proposes that the motion responsible for pattern formation occurs during the freezing of ground (differential frost heave) \cite{ar01, ar02}. This model yields a plausible mechanism for patterned ground but has yet to be experimentally confirmed, and does not explain certain observations such as the soil convection observed to occur in later summer \cite{br1}. The soil convection can potentially be explained by a phenomenon similar to the one that occurs in the formation of surface seals. The ice melts in late summer and at the same time evaporation takes place from the surface of the soil. This causes an unstable density profile to develop that is in principle large enough to initiate wholesale soil convection. This convection has been observed in fields and is thought to contribute to the patterns that form in Arctic region \cite{br1, ara0}. It may be noted that similar patterns have been detected on Mars, suggesting that in the past the temperature on Mars may have been large enough for the ice to melt and soil convection to take place \cite{mars}. In this paper we will use tensor notations for performing the stability analysis \cite{ti}. We will also use a new approach based on Green's functions to perform stability analysis. As will be evident that the stability analysis of a ternary mixture is very complicated, and it would not be possible to perform it using the conventional methods that are usually employed for stability analysis. \section{Ternary Mixture} In this section we will first analyse the evaporation of a mixture of soil particles and water. As the liquid evaporates, both the temperature and the concentration change at the surface. This causes an instability to occur which drives convection. In order to analyse the occurrence of this stability, we first observe that the density of the liquid depends on both the temperature of this mixture and concentration of particles in water. However, in any real situation there will also be various solute impurities in the soil. Hence, we analyze a ternary mixture of water, soil particles and dissolved solutes. Thus, we can write, \begin{equation} \rho = \rho_0 [ 1 - \alpha_c (C- C_0) + \alpha_c' (C'- C'_0) + \alpha_t ( T-T_0)], \end{equation} where $C$ is the concentration of soil particles and $C'$ is the concentration of solute impurities. The coefficients $\alpha_c$, $\alpha_c'$ and $\alpha_t$ are the particle, solutal and thermal expansion coefficients, respectively, taken to be constants. We now let $C \to -C \Delta C + C_0, C' \to -C' \Delta C' + C'_0$ and $ T \to -T \Delta T + T_0 $. We also define $B_c = \alpha_c \Delta C, B'_c = \alpha_c' \Delta C' $ and $B_t = \alpha_t \Delta T $ and so we get, $ \rho = \rho_0 [ 1 + B_c C + B_c' C' + B_t T ] $. Our system is described by a divergenceless vector field, which represents the fluid velocity, $ \partial^i v_i =0 $. In order to write the momentum balance equation, it is useful to define a substantive derivative as follows \begin{equation} D v_i = \partial_t v_i + v^j \partial_j v_i, \end{equation} where $ \partial_t = \partial/\partial t, \,\, \partial_i = \partial/\partial x^i, $ and $ \partial^i \partial_i = \partial^2 $. In continuum mechanics, this describes describes the time rate of change of some physical quantity for a material element subjected to a space-and-time-dependent velocity field. Now we can write the momentum balance equation as \begin{equation} \rho_0 D v_i = - \partial_i p + \partial^j \mu [1- C/C_p]^{-2} ( \partial_i v_j + \partial_j v_i) - \rho g \lambda_i. \end{equation} Here $\rho$ is the density, $g$ is the acceleration due to gravity, $\mu$ is a constant viscosity and $C_p$ is a maximum close packing concentration (shrinkage limit) \cite{wb1}. The factor $[1- C/C_p]^{-2}$ denotes the dependence of viscosity on concentration. Along with this equation there are also the following equations, \begin{eqnarray} DT &=& \kappa \partial^2 T, \nonumber \\ DC &=& \partial^i d_{11} [1- C/C_p]^{-2}\partial_i C + d_{12} \partial^2 C', \nonumber \\ DC' &=& d_{21}\partial^2 C + d_{22} \partial^2 C', \end{eqnarray} where $\kappa$ is the thermal diffusivity of the fluid, $d_{11}$ is the soil diffusion coefficient, $d_{22}$ is the solute diffusion coefficient, and $d_{21}, d_{12}$ are cross diffusion coefficients. We have neglected both the Dufour effect and Soret effect in the temperature equation as they are very weak relative to other effects. Here the thermal diffusivity of the fluid, solute diffusion coefficient and the cross diffusion coefficients are constants. As we will be analysing the this system far away from $C_p$, we will neglect the factor $C/C_p$ and so in this limit even soil diffusion coefficient and the viscosity is a constant. We will also use the Boussinesq approximation and set density constant in all terms except $\rho g \lambda_i$. We can use the natural length scale $l$, which is the depth of the liquid, and thermal diffusivity $k$ to non-dimensional these equations. To do so we will transform each quantity as $x^i \to l x^i, \partial_i \to l^{-1} \partial_i, v_i \to k l^{-1} v_i, t \to l^2 k^{-1} t,$ and $p \to p \mu k l^{-2} + \rho_0 g (l- \lambda^i r_i) $. The new velocity field again remains divergenceless, $ \partial^i v_i = 0 $. The non-dimensional form of the momentum conservation equation and the continuity equation for temperature and concentration for constant viscosity and diffusion coefficients becomes \begin{eqnarray} {Pr^{-1} D v_i } &=& - \partial_ i p + \partial^2 v_i+ R_c C \lambda_i + R_c' C' \lambda_i +R_t T \lambda_i , \nonumber \\ {D C }&=& Le^{-1}_{11} \partial^2 C + Le^{-1}_{12} \partial^2 C', \nonumber \\ {D C' }&=& Le^{-1}_{21} \partial^2 C' + Le^{-1}_{22} \partial^2 C' , \nonumber \\ {D T} &=& \partial^2 T, \end{eqnarray} where $ Pr = k \rho_0 \mu^{-1}$ is the Prandtl number, $R_c = \mu^{-1} k^{-1}B_c \rho_0 g l^3 $ and $R_c' = \mu^{-1} k^{-1}B_c' \rho_0 g l^3 $ are the concentration Rayleigh numbers, $R_t = \mu^{-1} k^{-1}B_t \rho_0 g l^3 $ is the temperature Rayleigh number, $Le_{11} = k d^{-1}_{11 }, Le_{22} = k d^{-1}_{22} $ are the Lewis numbers and $Le_{12} = k d^{-1}_{12}, Le_{21} = k d^{-1}_{21}$ are new Lewis number corresponding to cross diffusivity. We want to analyze the steady state version of these equations. In order to do so, we impose the following boundary conditions, $\lambda^i v_i = 0 $, $ \lambda^i \Lambda^j \partial_i v_j = 0 $, where $ \Lambda^i \lambda_i = 0$, and $C =C'= T = 1$, $C(1)= C'(1)= T(1) =0$, $ p(1) =0$. We also take $\lambda_i = (0, 0, 1)$ and $r_i = (x, y, z)$, thus our liquid is placed on a level plane. Under these boundary conditions we can obtain the steady state solutions. We will also add small perturbations to these steady state solutions. Now the steady state solutions with perturbations added to them is are given by \begin{eqnarray} v_i &=& u_i \nonumber \\ C &=& 1-\lambda^i r_i + \theta_c \nonumber \\ C' &=& 1-\lambda^i r_i + \theta_c' \nonumber \\ T &=& 1-\lambda^i r_i + \theta_t \nonumber \\ p &=& -\frac{ (R_c + R_c'+ R_t)}{2}|1-\lambda^i r_i|^2 + \phi. \end{eqnarray} Now we define $D^a_b = a^{-1} \partial_t - b^{-1} \partial^2$, and so we get \begin{eqnarray} D^{Pr}_1 u_i + \partial_i \phi - ( R_c \theta_c +R_c' \theta_c ' + R_t \theta_t ) \lambda_i&=&0, \nonumber \\ D^1_{Le11} \theta_c - Le_{12}^{-1} \partial^2\theta_c' - \lambda^i u_i &=&0, \nonumber \\ D^1_{Le22} \theta_c' - Le_{21}^{-1} \partial^2\theta_c - \lambda^i u_i &=&0, \nonumber \\ D^1_1 \theta_t - \lambda^i u_i &=&0. \end{eqnarray} We define an operator $E$ which takes the curl of any vector field $a^i$ twice, \begin{eqnarray} E a^i &=&\epsilon^{imn}\epsilon_{nkl} \partial_m \partial^k a^l \nonumber \\ &=& (\delta^i_{k} \delta_{l}^m - \delta^i_{l}\delta_{k}^m )\partial_m \partial^k a^l \nonumber \\ &=& \partial^i \partial^p a_p - \partial^2 a^i. \end{eqnarray} So, the gradient of a scalar field vanishes when its curl is taken twice, $ E \partial^i \phi = 0 $, because in the expression $ E \partial^i \phi = \epsilon^{imn}\epsilon_{nkl} \partial_m \partial^k \partial^l \phi $ there is a contraction between a pair of symmetric and antisymmetric tensor indices. We also have $E u_i = - \partial^2 u_i $, because $u_i$ is divergenceless, and $ \lambda^i (E \lambda_i \phi) = \lambda^i\partial_i \lambda^i \partial_j \phi - \partial^2 \phi = \tilde \partial^2 \phi $. Now acting on the momentum balance by $E$ and contracting it with $\lambda_i$, we get \begin{eqnarray} D^{Pr}_1 \partial^2 \lambda^i u_i - R_c \tilde \partial^2 \theta_c - R_c' \tilde \partial^2 \theta_c' - R_t \tilde \partial^2 \theta_t = 0. \end{eqnarray} Now we define the following differential operator \begin{eqnarray} \mathcal{D}_{0} &=& D^1_{Le 22} D^1_{Le 11} - Le^{-1}_{12}Le^{-1}_{21} \partial^4, \nonumber \\ \mathcal{D}_{1} &=& (D^1_{Le22 } - Le^{-1}_{12} \partial^2) , \nonumber \\ \mathcal{D}_{2} &=& (D^1_{Le11} - Le^{-1}_{21} \partial^2) , \nonumber \\ \end{eqnarray} and the following Green's function's \begin{eqnarray} \mathcal{D}_{0} G_c (r, r' ) &=& \delta^{3} (r-r'), \nonumber \\ D^1_1 G_t (r, r') &=& \delta^{3} (r-r'). \end{eqnarray} Now we have \begin{eqnarray} \mathcal{D}_{0} \theta_c &=& \mathcal{D}_{1} \lambda^i u_i, \nonumber \\ \mathcal{D}_{0} \theta_c' &=& \mathcal{D}_{2} \lambda^i u_i, \nonumber \\ D^1_1 \theta_t&=&\lambda^i u_i. \end{eqnarray} So, we have \begin{eqnarray} \theta_c (r) &=& \int d^3 r' G_c (r, r' ) \mathcal{D}_{1} \lambda^i u_i (r'), \nonumber \\ \theta_c' (r) &=& \int d^3 r' G_c (r, r' ) \mathcal{D}_{2} \lambda^i u_i (r'), \nonumber \\ \theta_t (r)&=&\int d^3 r' G_t (r, r' )\lambda^i u_i (r). \end{eqnarray} because \begin{eqnarray} \mathcal{D}_{0} \theta_c (r) &=&\int dr' \mathcal{D}_{0}G_c (r, r')\mathcal{D}_{1} \lambda^i u_i (r'), \nonumber \\ &=& \int d^3 r' \delta^3(r-r')\mathcal{D}_{1} \lambda^i u_i (r') \nonumber \\ &=& \mathcal{D}_{1}\lambda^i u_i (r), \nonumber \\ \mathcal{D}_{0} \theta_c (r) &=&\int dr'\mathcal{D}_{0} G_c (r, r')\mathcal{D}_{2}\lambda^i u_i (r'), \nonumber \\ &=& \int d^3 r' \delta^3(r-r')\mathcal{D}_{2} \lambda^i u_i (r') \nonumber \\ &=& \mathcal{D}_{2}\lambda^i u_i (r), \nonumber \\ { D^1_1 }\theta_t (r) &=&\int dr' D^1_1 G_t(r, r')\lambda^i u_i (r') \nonumber \\ &=& \int d^3 r' \delta^3(r-r')\lambda^i u_i (r') \nonumber \\ &=& \lambda^i u_i (r). \end{eqnarray} Thus, we can write, \begin{eqnarray} D^{Pr}_1 \partial^2 \lambda^i u_i (r) &=& R_c \tilde \partial^2 \int d^3 r' G_c (r, r' ) \mathcal{D}_{1} \lambda^i u_i (r') \nonumber \\ && + R_c' \tilde \partial^2 \int d^3 r' G_c (r, r' ) \mathcal{D}_{2} \lambda^i u_i (r') \nonumber \\ && + R_t \tilde \partial^2 \int d^3 r' G_t (r, r' )\lambda^i u_i (r'). \end{eqnarray} Multiplying by $\mathcal{D}_0$ and $D^1_1$, we get \begin{eqnarray} \mathcal{D}_0 D^1_1 D^{Pr}_1 \partial^2 \lambda^i u_i &=& R_c \tilde \partial^2 D^1_1 \mathcal{D}_{1} \lambda^i u_i + R_c' \tilde \partial^2 D^1_1 \mathcal{D}_{2} \lambda^i u_i \nonumber \\ && + R_t \tilde \partial^2 \mathcal{D}_0 \lambda^i u_i . \end{eqnarray} Now using the boundary conditions that the even derivatives of $\lambda^i \partial_i$ vanishes on $\lambda^i u_i $, we write the solution as \cite{ti}, \begin{equation} \lambda^i u_i = A \sin n\pi \, exp [i(k_x x + k_y) + \sigma t], \end{equation} becomes and we also define \begin{eqnarray} k^2 = k_x^2 + k^2_y, && k_n^2 = k^2 + n^2 \pi^2. \end{eqnarray} Thus the { characteristic} equation is \begin{equation} C(k_m, k, \sigma ) =0, \end{equation} where \begin{eqnarray} C(k_m, k, \sigma )& =& ( (\sigma + k_n^2 Le^{-1}_{22} ) (\sigma + k_n^2 Le^{-1}_{11} ) - Le^{-1}_{12}Le^{-1}_{21} k_n^4 )\nonumber \\ && \times (\sigma + k_n^2 ) (\sigma Pr^{-1} + k_n^2 ) k_n^2 \nonumber \\ &&- R_c (\sigma + k_n^2 ) ((\sigma + k_n^2 Le^{-1}_{22} ) - Le^{-1}_{12} k_n^2) k^2 \nonumber \\ && - R_c' (\sigma + k_n^2 ) ((\sigma + k_n^2 Le^{-1}_{11} ) - Le^{-1}_{12} k_n^2) k^2\nonumber \\ && - R_t ( (\sigma + k_n^2 Le^{-1}_{22} ) (\sigma + k_n^2 Le^{-1}_{11} ) \nonumber \\ && - Le^{-1}_{12}Le^{-1}_{21} k_n^4 ) k^2. \end{eqnarray} Assuming the principle of exchange of stabilities critical behavior is obtained by setting $\sigma =0$ \cite{ti}, we have \begin{equation} C(k_m, k, 0 ) =0, \end{equation} where \begin{eqnarray} C(k_m, k, 0)&=& ( Le^{-1}_{22} Le^{-1}_{11} - Le^{-1}_{12}Le^{-1}_{21} ) k_n^6 k^{-2} \nonumber \\ &&- R_c ( Le^{-1}_{22} - Le^{-1}_{12} ) \nonumber \\ && - R_c' ( Le^{-1}_{11} - Le^{-1}_{12} ) \nonumber \\ && - R_t ( Le^{-1}_{22} Le^{-1}_{11} - Le^{-1}_{12}Le^{-1}_{21} ). \end{eqnarray} Now we can define a effective Rayleigh number as \begin{eqnarray} R &=& [R_c ( Le^{-1}_{22} - Le^{-1}_{12} )+ R_c' ( Le^{-1}_{11} - Le^{-1}_{12} )\nonumber \\ &&+ R_t ( Le^{-1}_{22} Le^{-1}_{11} - Le^{-1}_{12}Le^{-1}_{21} ) ]\nonumber \\ &&\times ( Le^{-1}_{22} Le^{-1}_{11} - Le^{-1}_{12}Le^{-1}_{21} )^{-1}. \end{eqnarray} Hence, we can write $R = k_n^6 k^{-2}$. The lowest value is for $n =1$ and the instability starts at \begin{equation} \frac{\partial R}{\partial k^2} = 0. \end{equation} This gives us $ k^2 = \pi^2/2 $, and the corresponding value of $ R$ will be given by $ R= 657.5 \sim 10^3. $ This is when the instability will start. This convection in ternary mixtures can be used to analyse interesting geological phenomenon. This is because soils can exhibit semi-permeability and osmosis through these semipermeable soils can be driven by gradient in salt concentration \cite{os1, os2, os0}. In fact, an excess pressure has been observed to exist during the diffusion of salty water through certain soils \cite{os4, os5}. Thus, it will be interesting to analyse this phenomenon using stability analysis for ternary systems. In the limit when there is no cross diffusion, $Le^{-1}_{12} = Le^{-1}_{21} =0$, we have the expected result $R = R_c L_{11} + R_c' L_{22} + R_t$. Thus, in this limit, the temperature Rayleigh number and all the concentration Rayleigh multiplied by there respective Lewis numbers add up to give the effective Rayleigh number. So, the effective Rayleigh number is enhanced by the existence of a salt gradient. This, also implies that convection can start much earlier in salty water than pure water. Thus, the evaporation of salty water at the surface of soils can set up interesting instabilities, governed by the ternary characteristic equation. \section{Binary Mixture} In this section we will neglect the effect due to impurities. However, we are not able to directly set $Le^{-1}_{22} = Le^{-1}_{21} = Le^{-1}_{12} =0$, in the effective Rayleigh number. This is because the Green's function used to derive this effective Rayleigh number is not well defined for these values. This is because the differential operator $\mathcal{D}_0$ has a zero eigenvalue for these values of the Lewis numbers. So, its inverse does not exist. Thus, we can not set $Le^{-1}_{22} = Le^{-1}_{21} = Le^{-1}_{12} =0$, in the effective Rayleigh number obtained by using this Green's function. To obtain a correct result we will have to repeat the above analysis for the binary solutions. However, it may be noted that all the analysis of the previous section, except the derivation of the Green's function remains, remains well defined, if we set $Le^{-1}_{22} = Le^{-1}_{21} = Le^{-1}_{12} =0$. So, now we set, $Le_{11} = Le, Le^{-1}_{22} = Le^{-1}_{21} = Le^{-1}_{12} =0, \rho = \rho_0 [1 + B_c C + B_t T]$, and use the following equations \begin{eqnarray} {Pr^{-1} D v_i } &=& - \partial_ i p + \partial^2 v_i+ R_c C \lambda_i + R_t T \lambda_i , \nonumber \\ {D C }&=& Le^{-1} \partial^2 C , \nonumber \\ {D T} &=& \partial^2 T. \end{eqnarray} If we repeat the above analysis, we will obtain the following perturbative equations \begin{eqnarray} D^{Pr} \partial^2 \lambda^iu_i - R_c \tilde \partial^2 \theta_c - R_t \tilde \partial^2 \theta_t &=&0, \nonumber \\ D^1_{Le} \theta_c - \lambda^iu_i &=& 0, \nonumber \\ D^1_1 \theta_t - \lambda^iu_i &=& 0. \nonumber \\ \end{eqnarray} Now we again define the following Green's functions \begin{eqnarray} D^1_{Le} G (r, r')&=& \delta^3 (r-r'), \nonumber \\ D^1_{1} G_t (r, r')&=& \delta^3 (r-r'). \end{eqnarray} Thus, we can write \begin{eqnarray} \theta_c (r) &=& \int d^3 r' G (r,r') \lambda^i u_i (r'), \nonumber \\ \theta_t (r) &=& \int d^3 r' G_t (r,r') \lambda^i u_i (r'), \end{eqnarray} because \begin{eqnarray} D^1_{Le} \theta_c (r) &=& \int d^3 r'D^1_{Le} G (r,r') \lambda^i u_i (r') \nonumber \\ &=& \int d^3 r' \delta (r-r') \lambda^i u_i (r') \nonumber \\ &=& \lambda^i u_i (r), \nonumber \\ D^1_{1} \theta_t(r) &=& \int d^3 r'D^1_{1} G_t (r,r') \lambda^i u_i (r') \nonumber \\ &=& \int d^3 r' \delta (r-r') \lambda^i u_i (r') \nonumber \\ &=& \lambda^i u_i (r). \end{eqnarray} We can write \begin{eqnarray} D^{Pr} \partial^2 \lambda^iu_i (r) &=& R_c \tilde \partial^2 \int d^3 r' G(r,r') \lambda^i u_i (r') \nonumber \\ && + R_t \tilde \partial^2 \int d^3 r' G_t (r,r') \lambda^i u_i (r'). \end{eqnarray} Now acting on this equation by $D^1_1$ and $D^1_{Le}$, we get \begin{equation} D^1_1 D^1_{Le} D^{Pr}_1 \partial^2 \lambda^iu_i = R_c \tilde \partial^2 D^1_1 \lambda^i u_i + R_t \tilde \partial^2 D^1_{Le} \lambda^i u_i (r'). \end{equation} So, again using the boundary conditions that the even derivatives of $\lambda^i \partial_i$ vanishes on $\lambda^i u_i $, we write the solution as \cite{ti} \begin{equation} \lambda^i u_i = A \sin n\pi \, exp [i(k_x x + k_y y) + \sigma t], \end{equation} Thus, we get \begin{equation} C(k_n, k, \sigma) =0, \end{equation} where \begin{eqnarray} C(k_n, k, \sigma) &=& (\sigma + k_n^2 Le^{-1} ) (Pr^{-1}\sigma + k_n^2 ) (\sigma + k_n^2 ) k_n^2 \nonumber \\ & & - (\sigma + k_n^2 Le^{-1} ) k^2 R_t - (\sigma + k_n^2 ) k^2 R_c. \end{eqnarray} So, in the case of temperature dependence critical behavior is obtained by setting $\sigma =0$, we have \cite{ti} \begin{equation} C(k_m, k, 0 ) =0, \end{equation} where \begin{eqnarray} C(k_n, k, 0) &=& k_n^6 k^{-2}Le^{-1} - Le^{-1} R_t - R_c. \end{eqnarray} Now here the effective Rayleigh number is \begin{equation} R = R_t + Le R_c \end{equation} Hence, we can write $ R = k_n^6 k^{-2}. $ The lowest value is for $n =1$ and the instability starts at \begin{equation} \frac{\partial R}{\partial k^2} = 0. \end{equation} This gives us $ k^2 = \pi^2/2 $, and the corresponding value of $ R$ will be again be given by $ R= 657.5 \sim 10^3 $. Now the actual value for a binary mixture of bentonite and water are of the order, $ \rho_0 \sim 10^3, \mu \sim 10^{-3}, \kappa \sim 10^{-7}$. Also the depth of the liquid $ l \sim 10^{-2}$ \cite{ee} and $B_c \sim 1, B_t \sim 10^{-1}$ \cite{ee11}, so we have $R_t \sim 10^3$, and $R_c \sim 10^6$. However, in the stability analysis the product of $R_c$ and $Le$ contributes to the occurrence of the instability and $Le \sim 10^2$. So, the contribution from the concentration gradient $R_c Le \sim 10^8$ is much more than temperature gradient. The Marangoni effect is also measured by the Marangoni number $M\sim 10^3$ \cite{ee22}. Hence, the concentration gradient is the most dominant factor for convection to occur. The calculated value of the effective Rayleigh number due to the concentration gradient is much larger than the theoretical limit for the convection to occur. Hence, we hypothesize that it is responsible for the instability in the drying of a binary mixture of bentonite and water. It has been observed that in a binary mixture of bentonite and water initially the convection takes place on the surface, but, eventually a layer forms on the surface and this inhibited further convection from taking place on the surface \cite{ee}. However, the convection continues in the lower depths of the soil. This observation rules out the possibility that this convection is due to the Marangoni effect. This is because in that case it would be surface driven phenomenon and would not continue at depth after halting at the surface. It will be interesting to explore the possibility that this behavior is due to the fact that in a soil the diffusivity and viscosity are strong functions of particle concentration. That is, the effective viscosity and diffusivity of a soil slurry increase dramatically with increasing particle concentration, and thus at sufficiently high particle concentrations the effective Rayleigh number will be reduced to below the critical value necessary for convection. \section{Conclusion} In this paper we have analysed the drying of a ternary mixture of soil particles and water mixed with impurities. We have demonstrated that the concentration gradient that forms during the process of evaporation of this mixture is the main driving force for the occurrence of a Rayleigh-Benard instability. This instability causes convection to take place, which in turn may have significant implications for the drying of soil crusts and the formation of patterned ground. In the ternary case the use of Green's functions allowed for a straight-forward calculation of the characteristic equation. For the binary case a separate calculation was required. This is because the differential operator that was inverted using the Green's function in the ternary case, yields a zero eigenvalue in the binary case. Thus, it cannot be inverted and the ternary Green's function becomes ill defined for the binary case. As noted in the introduction, soil loss in agricultural areas depends critically on the formation of surface seals and crusts and the subsequent cracking of these crusts during drying of the ground. It is possible that the convection patterns give rise to weak points, which on drying gives rise to cracks \cite{ee}. Furthermore, patterns observed during the freezing and thawing of ground around the Arctic circle have also been explained as owing to convection of the soil. In this paper it was proposed that evaporation from the surface of thawing ground causes an unstable density profile to develop potentially leading to convection. The occurrence of similar patterns on Mars suggests that in the past the temperature on Mars may have been large enough for the ice to melt and evaporation to take place. It may be noted that it has been observed that drying on slanting slopes gives rise to rolls and drying on flat surface gives rise to hexagonal structures \cite{ara0}. It will be interesting to perform stability analysis for these specific physical situations and see if the formation of these particular patterns can be explained.
\section{Introduction} Observations of the local universe support the conclusions drawn by the hierarchical theory of structure formation that the majority of galaxies reside in groups, dynamically-bound systems which span a wide range of properties \citep{tully87}. The intra-group medium (IGM) contained within these groups likely contains a significant fraction of the baryonic content of the universe \citep{Fuk98}. Observations measuring the baryon content of the local universe account for approximately a third of the baryon density observed at high redshift ($z = 2-4$) in the form of stars, cold gas, and hot X-ray-emitting gas \citep{Fuk04,Stocke04}. Numerical simulations predict a significant fraction of the ``missing'' baryons, around $40\%$ of the total baryonic content of the local universe, may be contained in the warm-hot intergalactic medium (WHIM) with a temperature range of $T \sim 10^5-10^7$~K \citep{Cen06,Dave01}. The primary observational tool for studying the WHIM is currently absorption-line spectroscopy of low-redshift quasars \citep{Nara10}. These observations are limited to groups falling along the line of sight of a sufficiently bright quasar, and density measurements depend on estimates of the systems' spatial extent, metallicity and ionizing fraction, yielding total IGM density measurements on the order $n \sim 10^{-4}-10^{-5}$~cm$^{-3}$ \citep{Pisano04}. X-ray measurements are inherently limited to higher temperature groups, generally containing at least one early-type galaxy. A dynamical mass can be estimated by making an assumption about the geometrical distribution of X-ray emitting gas in the group. However, measurements of X-ray surface brightness drop below measurable levels well within the virial radius, requiring additional assumptions in order to extrapolate the total mass estimate of the group \citep{Mulchaey00}. Observations of bent-double radio sources in galaxy groups serve as a valuable method of measuring IGM densities, and recently this method has been used to find total IGM densities in groups of $n \sim 2\times10^{-4} - 3\times10^{-3}$~cm$^{-3}$ \citep{Free08,Free10,Free11}, higher than the density found from absorption-line measurements. These measurements rely on a set of assumed physical parameters including viewing angle, AGN proper motion, and kinematic luminosity. These assumptions are largely independent of those required for UV and X-ray density estimates, making this analysis complementary to existing methods. Bent-double radio sources also probe the density of the entire IGM, rather than just one temperature phase. Ram pressure resulting from the proper motion of these double-lobed radio sources through the IGM sweeps back the bipolar jets, producing the distinctive bent-double radio tails noted in many radio observations \citep{Miley72}. Prior to \citet{Burns87}, it was believed that the conditions on ambient IGM density and galaxy velocities necessary to produce these bent-double radio tails could be found only in large, rich clusters of galaxies. However, surveys have found a significant number of bent-double sources in lower mass galaxy groups \citep{Venk94,Doe95,Blanton01}. \citet{Ekers78} was among the first to identify bent-double radio sources as a possible density probe of the IGM, observing IGM densities in the range $n \sim 3 \times 10^{-4}-6 \times 10^{-3}$~cm$^{-3}$ in the galaxy groups NGC~6109 and NGC~6137. These early estimates have been corroborated by more recent observations of head-tail sources in galaxy groups and poor clusters \citep{Free08,Free11}. If these numbers are representative of all galaxy groups, a significant fraction of the local universe's baryon content would reside within the IGM. Analytic modeling of bent-double radio sources has found that the radius of curvature at the most bent part of the jet can be described by an equation of the form \begin{eqnarray} \frac{R}{h} = \frac{P_{\rm jet}}{P_{\rm ram}} \label{eqn:r_over_h} \end{eqnarray} \noindent which balances internal and external pressure gradients as a ratio of the radius of curvature $R$ and the scale height $h$ across which the pressure difference acts \citep{Begel79,Jones79,Burns80,Odea85}. In \citet{Begel79}, $h$ is taken to be the diameter of the jet, whereas in Jones \& Owen, $h$ is the scale height of the ISM within the host galaxy. We use the diameter of the jet for $h$. We use this relationship as an analytic model to compare to our simulation results. We carry out a series of numerical simulations of bent-double radio sources to determine values for $R$ and $h$ in terms of initial model parameters. In \S 2, we describe the numerical methods and initial conditions used in our simulations. In \S 3, we describe the results of our simulations and our methods for measuring $R$ and $h$. In \S 4, we use our simulations to quantify the errors on density estimates from observations of bent-double radio sources. We also develop a formula to estimate the kinetic luminosity of observed jets, and we make predictions for the X-ray detectability of radio sources. In \S 5, we summarize our results. \section{Technical description} \subsection{Code description} Simulations are carried out using the FLASH 2.4 hydrodynamics code \citep{Fryxell00}, which is a modular, block-structure adaptive mesh code. It solves the Riemann problem on a three-dimensional Cartesian grid using the piecewise-parabolic method. Gas is modeled as having a uniform adiabatic index of $\gamma = 5/3$. \subsection{The Jet Nozzle} In order to simulate the injection of collimated, supersonic jets into the grid, we employ a numerical ``nozzle'', as first developed and described in \citet{Heinz06}: an internal inflow boundary of cylindrical shape placed at the location of the AGN, injecting fluid with a prescribed energy, mass, and momentum flux to match the parameters we choose for the jet. For reasons of numerical stability, we impose a slow lateral outflow with low mass flux in order to avoid complete evacuation of zones immediately adjacent to the nozzle due to the large velocity divergence at the nozzle. The injection of energy and mass due to this correction is negligible. We model unresolved dynamical instabilities near the base of the jet by imposing a random-walk jitter on jet axis confined to a $5\degr$ half-opening angle. This is necessary to model the `dentist's drill' effect of \citet{Scheuer82}. The time for the jet to change direction is slow compared to time for jet material to reach the end of the jet, so the jitter does not significantly change the bending of the jet, aside from symmetry breaking. It does, however, change the shape of the initial cocoon created around the jet (see section~\ref{sect:results}) by spreading energy over a wider average angle. We chose to inject the jet at an internal Mach number of 10 in most simulations. For computational feasibility, we chose a jet velocity of $v_{\rm jet} = 3 \times 10^9$~cm/s for most simulations. For a typical case (i.e. model .25E), this results in a jet density of $n_{\rm jet} = 1.82\times10^{-5}$~cm$^{-3}$, pressure of $p_{\rm jet} = 1.64\times10^{-12}$~erg$/$cm$^{3}$ and temperature of $T_{\rm jet} = 6.53\times10^8$~K. The jet is turned on initially and continues to inject material for the entire length of the simulation. \subsection{Initial Conditions} Our setup is similar to that used for X-ray binary jets in \citet{Yoon11}. We place the jet nozzle in a moving medium inside a simulation box large enough that the boundaries never affect the simulation ($2.8$~Mpc on a side). We keep the location of the AGN fixed in space, letting the IGM stream by at velocity $-v_{\rm gal}$ perpendicular to the jet axis. We can vary the velocity and density of the IGM, as well as the luminosity, velocity and internal Mach number of the jet. In all cases the IGM pressure is set to $2.76 \times 10^{-13}$~erg/cm$^3$, giving a typical IGM sound speed of $166$~km/s and temperature of $2\times10^6$~K. A list of parameters for all simulations is given in table~\ref{table:model_data}. The simulations were carried out on a staggered mesh grid as described in \citet{Yoon11} in order to capture the large dynamic range required, and ensure that the nozzle diameter is resolved by 12 grid cells. For our normal scaling, the nozzle diameter is 2~kpc, with a maximum resolution for the standard model of about 0.175~kpc near the jet nozzle. The nozzle diameter for each simulation, d$_{\rm j}$, is listed in table~\ref{table:model_data}. In all cases, the maximum resolution is set such that the nozzle is resolved by 12 grid cells. Short duration test simulations at higher resolution were carried out and produced similar results. One simulations was for a conical jet with an initial opening angle, rather than a purely parallel inflow. This model, .015E\_theta20, has identical parameters to model .015E, except that the nozzle is 4 times smaller (and the maximum resolution 4 times greater) and the inflowing material is spread over a $20\degr$ half-opening angle. This model was run as a direct comparison to model .015E, based on the analytic model in section~\ref{sect:analytic}. We use cylindrical (parallel) jet injection for our other models because this allows us to use a much lower resolution. \begin{table*} \centering \begin{minipage}{140mm} \caption{Model Data} \begin{tabular}{@{}lllllllll@{}} \hline Model & $L_{\rm jet}$ & $n_{\rm IGM}$ & $v_{\rm gal}$ & $v_{\rm jet}/c$ & M & R & h & d$_{\rm j}^a$ \\ ~ & $(10^{44}$~erg/s) & (cm$^{-3}$) & (km/s) & ~ & ~ & (kpc) & (kpc) & (kpc) \\ \hline .015E\_theta20 & $0.015625$ & $1\times10^{-3}$ & $1000$ & $0.1$ & $10$ & $8.6\pm0.8$ & $0.44\pm.02$ & 0.125$^b$ \\ .015E & $0.015625$ & $1\times10^{-3}$ & $1000$ & $0.1$ & $10$ & $9.4\pm1.4$ & $0.49\pm.01$ & 0.5 \\ .062E & $0.0625$ & $1\times10^{-3}$ & $1000$ & $0.1$ & $10$ & $18.4\pm5.2$ & $0.91\pm.03$ & 1.0 \\ .25E & $0.25$ & $1\times10^{-3}$ & $1000$ & $0.1$ & $10$ & $35.5\pm7.9$ & $2.06\pm.07$ & 2.0 \\ 1E & $1.0$ & $1\times10^{-3}$ & $1000$ & $0.1$ & $10$ & $76.4\pm15.3$ & $3.87\pm.14$ & 4.0 \\ .062E\_.5vel & $0.0625$ & $1\times10^{-3}$ & $500$ & $0.1$ & $10$ & $35.4\pm6.9$ & $2.05\pm.06$ & 2.0 \\ .015E\_.25vel & $0.015625$ & $1\times10^{-3}$ & $250$ & $0.1$ & $10$ & $38.9\pm10.8$ & $1.94\pm.07$ & 2.0 \\ 1E\_4n & $1.0$ & $4\times10^{-3}$ & $1000$ & $0.1$ & $10$ & $35.5\pm12.2$ & $2.09\pm.10$ & 2.0 \\ .062E\_.25jvel & $0.0625$ & $1\times10^{-3}$ & $1000$ & $0.025$ & $10$ & $30.4\pm3.3$ & $2.41\pm.14$ & 2.0 \\ .25E\_4M & $0.25$ & $1\times10^{-3}$ & $1000$ & $0.1$ & $40$ & $40.1\pm6.0$ & $1.89\pm.05$ & 2.0 \\ \hline \label{table:model_data} \end{tabular} $^a$ Nozzle diameter at injection $^b$ This simulations has a half-opening angle of $20\degr$ at injection, rather than a parallel inflow \end{minipage} \end{table*} \section{Simulation Results } \label{sect:results} When the jet initially turns on, it drives a shock into the IGM, creating a rapidly expanding cocoon. The expansion velocity of the cocoon is initially faster than $v_{\rm gal}$, so the AGN remains inside the cocoon and the jet stays fairly straight. As the cocoon expands it decelerates, but $v_{\rm gal}$ remains constant. Eventually, the AGN moves outside the initial cocoon and a bow shock is created around the AGN jets. The pressure gradient from this shock bends the jets backwards, creating the characteristic curved structure. At some point along the jet, the flow becomes unstable and breaks up, creating bright radio lobes at the ends of the jets. The swept-back jet material creates a long tail of diffuse jet material reaching back to the cocoon created around the initial position of the AGN. We create synthetic radio images by assuming that the jet material, followed with a tracer fluid, consists of a hot plasma of relativistic particles with magnetic fields in equipartition with thermal pressure. Energy lose from synchrotron cooling is not included. The left panel of Fig.~\ref{fig:radio_example} shows a synthetic radio image for simulation .25E at 50~Myr. In this image the AGN is moving relative to the IGM from right to left. The narrow curved jets and bright radio lobes can be clearly seen. The tail and cocoon produce diffuse radio emission with a surface brightness 1 to 2 orders of magnitude fainter than the jet and lobes. This extended emission is typically not seen in bent-double radio sources in jets. It is possible that this emission is resolved out in existing observations or that synchrotron cooling, which is not included in making our images, makes the emission at $1.4$~GHz too faint to observe. Observations at lower resolution and/or lower frequencies may reveal the extended tail. It is possible that at least part of the tail is seen in source S7 in \citet{Free11}, which appears to have extremely thick jets relative to the radius of curvature. However, this is not the only explanation for this source's appearance (see section~\ref{sect:viewing_angle}). The length of the tail is $v_{\rm gal} \times t_{\rm AGN}$, where $t_{\rm AGN}$ is the amount of time the AGN has been active. If there is an estimate for the AGN host galaxy velocity, finding the length of the tail would allow the amount of time the AGN has been active to be determined. Note that in Fig.~\ref{fig:radio_example} the length of the tail (left to right) is less than the width of the tail (top to bottom), because the jets initially expand outwards faster than $v_{\rm gal}$. If the AGN remains active, the tail will continue to lengthen and eventually become longer than it is wide. \subsection{Radius of Curvature \label{sect:measure_r}} The right panel of Fig.~\ref{fig:radio_example} shows the same image with a circle overlaid with $R = 37$~kpc, the best fit radius of curvature for this image. The circle traces both the upper and lower jet and passes through the bright radio lobes where each jet breaks up. Note that the upper lobe is significantly brighter than the lower lobe at this time. The brightness of the lobes and the curvature of the two jets varies with time due to instabilities in the jet propagation and the small random changes in jet direction that we introduce. The radius of curvature is determined by an automated fitting procedure. We start with a synthetic radio image and consider only the region $28$~kpc above and below of the AGN, the approximate extent of the jets. Because the upper and lower jets and lobes can differ significantly in brightness, we find the maximum brightness of any point in each half of the image and exclude points with less than $10\%$ of this brightness. For each slice along the jet, we then determine the horizontal location of the jet by taking an intensity weighted average of the radio emission in that slice. We then fit a circle to the jet locations weighted by the total radio intensity of the points in each slice. To characterize how much fluctuations in the jet affect the radius of curvature, we use the above procedure to measure the radius of curvature at a series of times (after the jet curvature is well established) and compare the results. Fig.~\ref{fig:curve_fits_.25E} plots measured values of $R$ for simulation .25E from $40$~Myr to $220$~Myr. The error bars represent the $1$-$\sigma$ error on the value of $R$ at each time, typically about $10\%$. The values of $R$ are very consistent, with a scatter of about $25\%$. The best-fit value of $R$ for simulation .25E is $35.5\pm7.9$~kpc, where the error takes into account both the error in individual fits and variations between different snapshots. The same procedure is applied to all of our simulations, with the results listed as $R$ in table~\ref{table:model_data}. \subsection{Jet Thickness} A similar process is used to find the average jet thickness, listed as $h$ in table~\ref{table:model_data}. The thickness typically has a very small variation, $\simeq 5\%$, and the accuracy is limited by the resolution of our simulations. Because we are interested in the ``real'' value of the jet thickness, we determine it using the raw simulation data rather than the synthetic radio data. We take a region of $\pm24$~kpc from the AGN and define points as being in the jet if at least $80\%$ of the material at that point was injected by the nozzle and it is moving with at least $80\%$ of the initial jet velocity. This excludes all material in the lobes and the cocoon around the jet. We then determine the jet thickness perpendicular to both the jet direction and the motion of the AGN, henceforth the `z' direction, as this is the front along which the ram pressure acts. For each slice through the jet we take the average thickness for points in the jet in the `z' direction. We then take the average of this thickness in all slices to get a value for $h$ at that time, with the standard deviation of the average being the uncertainty at that time. To get the values of $h$ in table~\ref{table:model_data}, we then find the best-fit value of the thickness for all times considered, and the error takes into account both the error at individual times and the variation with time. The overall result is that the radius measured at a particular time is accurate to within an error of 25\% of the `true' value for that particular combination of parameters. The thickness is more consistent, with typically only a $5\%$ variation. The variation in $R$ sets a lower limit on the accuracy of the IGM density estimated from observations of bent-double radio sources. \section{Discussion} \subsection{Analytic Fit for Radius of Curvature \label{sect:analytic}} For an AGN moving supersonically relative to the IGM, the jets will curve due to the ram pressure of the incoming IGM. At the point of maximum curvature, the ratio of the radius of curvature to the thickness of the jet will equal the ratio of the ram pressure of the jet to the external ram pressure \citep{Begel79,Jones79,Burns80}, i.e.: \begin{eqnarray} \frac{R}{h} = \frac{P_{\rm jet}}{\rho_{\rm IGM} v_{\rm gal}^2} \label{eqn:r_over_h} \end{eqnarray} \noindent The ram pressure of the jet will be \begin{eqnarray} P_{\rm jet} = \frac{L_{\rm jet}}{\frac{\pi}{4} h^2 v_{\rm jet}} \label{eqn:P_jet} \end{eqnarray} Rearranging eqn.~\ref{eqn:r_over_h} and eqn.~\ref{eqn:P_jet} gives a formula for $R$ in terms of $h$: \begin{eqnarray} R = \frac{L_{\rm jet}}{\frac{\pi}{4} h v_{\rm jet} \rho_{\rm IGM} v_{\rm gal}^2} \label{eqn:r_with_h} \end{eqnarray} \noindent However, the thickness of the jet is not constant and will be determined by the external pressure. Initially the lateral expansion of the jet will be ballistic because the component of the ram pressure perpendicular to the jet will be higher than the external pressure. However, as the thickness increases the internal pressure will drop until it is the same order as the cocoon (i.e. external) pressure. At this point, a recollimation shock will be driven into the jet, setting the thickness and providing an internal pressure that is equal to that of the cocoon. Therefore, the thickness will be set such that the external pressure balances the perpendicular ram pressure. For a jet with an initial half-opening angle of $\theta$, the average ram pressure of the jet perpendicular to direction of motion will be $P_{\rm jet,\perp} \simeq \frac{1}{2} P_{\rm jet} \sin^2 \theta$. Balancing this against the external pressure gives \begin{eqnarray} \rho_{\rm IGM} v_{\rm gal}^2 = \frac{1}{2} P_{\rm jet} \sin^2 \theta = \frac{1}{2} \frac{L_{\rm jet}}{\frac{\pi}{4} h^2 v_{\rm jet}} \sin^2 \theta \label{eqn:p_perp} \end{eqnarray} \noindent Solving for h, we find \begin{eqnarray} h = \frac{\sin \theta}{\sqrt 2} \left( \frac{L_{\rm jet}}{\frac{\pi}{4} v_{\rm jet} \rho_{\rm IGM} v_{\rm gal}^2} \right)^{1/2} \label{eqn:h} \end{eqnarray} \noindent Using this value for $h$ in eqn.~\ref{eqn:r_with_h} gives us \begin{eqnarray} R = \frac{2 h}{\sin^2 \theta} = \frac{\sqrt 2}{\sin \theta} \left( \frac{L_{\rm jet}}{\frac{\pi}{4} v_{\rm jet} \rho_{\rm IGM} v_{\rm gal}^2} \right)^{1/2} \label{eqn:R} \end{eqnarray} For our simulations with a cylindrical jet, we fix the nozzle size to h set by \begin{eqnarray} h &=& 4\textrm{~kpc} \times \left(\frac{L_{\rm 44}}{n_{\rm -3} v_{\rm gal,1000}^2 v_{\rm jet,0.1}}\right)^{1/2} \label{eqn:h_for_parallel} \end{eqnarray} \noindent where $L_{\rm 44} = L_{\rm jet}/10^{44}$~erg/s, $n_{\rm -3} = n_{\rm IGM}/10^{-3}$~cm$^{-3}$, $v_{\rm gal,1000} = v_{\rm gal}/1000$~km/s, and $v_{\rm jet,0.1} = v_{\rm jet}/(0.1 c)$. This is the equivalent width of a jet with an initial opening angle of $\theta = 20\degr$. From eqn.~\ref{eqn:R}, this would predict a value of $h/R = 1/17$. We run one model, .015E\_theta20, with a small nozzle size and a $20\degr$ initial opening angle. The measured values of $R$ and $h$ in talbe~\ref{table:model_data} are $8.6\pm0.8$~kpc and $0.44\pm0.2$~kpc, respectively, very close to the prdicted values of $8.5$~kpc and $0.49$~kpc given by equations~\ref{eqn:R} and \ref{eqn:h}. The results are also very similar to model .015E, which has the same setup but with a cylindrical jet input and a $0.5$~kpc nozzle size. Using the measured values of $R$ and $h$ in table~\ref{table:model_data} for all of our parallel jet models, we find that the value of $h$ stays very close to the nozzle size and the ratio of $h/R$ is about $1/18$, close the our analytic estimate of $1/17$. Using the measured values of $R$ in table~\ref{table:model_data}, we find that the value of $R$ in terms of our model parameters is \begin{eqnarray} R &=& 5\textrm{~} \left(\frac{L_{\rm jet}}{\rho v_{\rm gal}^2 v_{\rm jet}}\right)^{1/2} \textrm{~cm} \nonumber \\ &=& 72 \textrm{~kpc} \times \left(\frac{L_{\rm 44}}{n_{\rm -3} v_{\rm gal,1000}^2 v_{\rm jet,0.1}}\right)^{1/2} \label{eqn:radius_fit} \end{eqnarray} Figure~\ref{fig:curve_vs_energy} plots the radius of curvature vs. jet luminosity for models .015E, .062E, .25E and 1E. These models differ only in jet luminosity, with all other parameters the same. Error bars represent the overall error in measuring $R$, as discussed in section~\ref{sect:measure_r}. The solid line is a plot of eqn.~\ref{eqn:radius_fit} for radius vs. luminosity and passes well within the error bars of all the four data points. Equation~\ref{eqn:R} can also be used to find the kinetic luminosity of observed bent-double radio sources based on their measured radius of curvature and jet pressure. In \citet{Free11}, $P_{\rm jet}$ is assumed to be the minimum synchrotron pressure \begin{eqnarray} P_{\rm jet} = P_{\rm min} = (2\pi)^{-3/7} \left( \frac{7}{12} \right) [c_{\rm 12} L_{\rm rad} (1+k) (\phi V)^{-1} ]^{4/7} \label{eqn:pmin} \end{eqnarray} \noindent where $c_{\rm 12}$ is a constant that depends on the spectral index and frequency cutoffs \citep{Pacholczyk70}, $k$ is the ratio of relativistic proton to relativistic electron energy, $\phi$ is the volume filling factor, $V$ is the source volume, and $L_{\rm rad}$ is the radio luminosity of the jet at the point where the pressure is measured. $P_{\rm jet}$ is measured in several slices along the jet, with the volume assumed to be proportional to the square of measured jet thickness, which is limited by the beam size of the observations. Therefore, $P_{\rm min} \propto h^{-8/7}$ but is independent of the radius of curvature $R$ and the length of the jet. Rearranging eqn.~\ref{eqn:R}, we find a jet luminosity of \begin{eqnarray} L_{\rm jet} &=& \frac{\pi}{8} R^2 \sin^2 \theta \rho_{\rm IGM} v_{\rm gal}^2 v_{\rm jet} \nonumber \\ &=& 3.73 R_{\rm kpc}^2 \left( \frac{h}{R} \right) n_{\rm IGM,-3} v_{\rm gal,1000}^2 \beta_{\rm jet} \times 10^{42} \textrm{~erg/s} \label{eqn:L_fit_dens} \end{eqnarray} \noindent where $R_{\rm kpc}$ is $R$ in kpc and $\beta_{\rm jet} = v_{\rm jet}/c$. In terms of pressure $L_{\rm jet}$ is \begin{eqnarray} L_{\rm jet} &=& \frac{\pi}{4} R^2 \left( \frac{h}{R} \right)^2 P_{\rm min} v_{\rm jet} \nonumber \\ &=& 2.24 h_{\rm kpc}^2 P_{\rm min,-11} \beta_{\rm jet} \times 10^{42} \textrm{~erg/s} \label{eqn:L_fit_pres} \end{eqnarray} \noindent where $h_{\rm kpc}$ is $h$ in kpc. The jet kinetic luminosity calculated using this formula for the sources and measured values of $R$, $h$ and the synchrotron pressure $P_{\rm min}$ in \citet{Free11} are listed in table~\ref{table:sources_emily}. The values in table~\ref{table:sources_emily} are upper limits on the luminosity made with the assumptions that the observed value of $h$ is the true jet thickness and that the jet velocity is $v_{\rm jet} = c$. If the jet is narrower or slower, the luminosities will be smaller. Note that the formula for $L_{\rm jet}$ is independent of the AGN velocity, so values can be found even for sources with unconstrained velocities. The sources are all of order $10^{45}$~erg/s and the variation between sources is smaller than the variation in the total radio power at $1440$~MHz ($L_{\rm 1440}$). $P_{\rm min}$ is proportional to $L_{\rm rad}^{4/7}$, which is the total radio emission at the point where $P_{\rm min}$ is measured. $L_{\rm rad}$ therefore depends on the $1440$~MHz emission at that point and the model of the synchrotron spectrum used, but is not directly dependent on the total $1440$~MHz emission of the entire source. $P_{\rm min}$ scales with jet thickness as $h^{-8/7}$, so $L_{\rm jet} \propto h^{6/7}$. Note that we assume the ram pressure of the jet, $P_{\rm jet}$, is accurately reflected by the minimum synchrotron pressure, $P_{\rm min}$, which is true only if the jet energy is dominated by relativistic electrons and magnetic fields. If the true jet ram pressure is higher, the estimates of both $n_{\rm IGM}$ and $L_{\rm jet}$ in table~\ref{table:sources_emily} will both be proportionally higher. Adding invisible components to the momentum flux of the jet only acts to increase the required IGM density. On a similar note, in our simulations very little external material becomes entrained in the jets. However, even if a significant amount of mass is entrained it will not change the momentum flux of the jet. Therefore, the radius of curvature and inferred IGM density should not affected. Also note that the equations in this section were derived assuming that the AGN host galaxy is moving supersonically relative to the IGM and that the jet velocity is supersonic relative to its internal sound speed. These relations are not expected to hold if galaxies or jets are subsonic. \begin{table*} \centering \begin{minipage}{180mm} \caption{Jet Kinetic Luminosity} \begin{tabular}{@{}llllllll@{}} \hline Source ID & $v_{\rm gal}^{a,b}$ & $P_{\rm min}^a$ & $n_{\rm IGM}^a$ & $h^a$ & $R_{\rm bend}^a$ & $L_{\rm 1440}^a$ & $L_{\rm jet}^c$ \\ ~ & (km/s) & $10^{-11}$~erg~cm$^{-3}$ & (cm$^{-3}$) & (kpc) & (kpc) & (W Hz$^{-1}$) & (erg/s) \\ \hline S1 & $430^{+170}_{\rm -35}$ & $0.9\pm0.2$ & $3\pm2 \times 10^{-3}$ & $23\pm1$ & $42\pm6$ & $1.6 \times 10^{25}$ & $1.1\pm0.2 \times 10^{45}$\\ S2 & $570\pm60^{d}$ & $0.6\pm0.2$ & $5\pm4 \times 10^{-4}$ & $30\pm4.5$ & $104\pm9$ & $1.88 \times 10^{25}$ & $1.2\pm0.5 \times 10^{45}$\\ S3 & $745^{+109}_{\rm -80}$ & $1.4\pm0.6$ & $2\pm1 \times 10^{-4}$ & $10\pm0.6$ & $141\pm19$ & $3 \times 10^{23}$ & $3.1\pm1.4 \times 10^{44}$\\ S4 & $950^{+210}_{\rm -140}$ & $1.7\pm0.3$ & $5\pm2 \times 10^{-4}$ & $22\pm0.7$ & $89\pm7$ & $9.5 \times 10^{24}$ & $1.5\pm0.3 \times 10^{45}$\\ S5 & Unconstrained & $0.4\pm0.1$ & $(70\pm21)/v^2$ & $38\pm3.8$ & $220\pm11$ & $6.4 \times 10^{24}$ & $1.3\pm0.4 \times 10^{45}$\\ S6 & Unconstrained & $0.6\pm0.1$ & $(156\pm48)/v^2$ & $19\pm3.1$ & $69\pm4$ & $9 \times 10^{23}$ & $4.8\pm1.4 \times 10^{44}$\\ S7 & $850^{+170}_{\rm -120}$ & $1.4\pm0.3$ & $2\pm1 \times 10^{-3}$ & $22\pm3.6$ & $18\pm4$ & $2 \times 10^{24}$ & $1.5\pm0.5 \times 10^{45}$\\ \hline \label{table:sources_emily} \end{tabular} $^a$~from~\citet{Free11}. $^b$~$v_{\rm gal}$ here is $\sqrt3$ times the group velocity dispersion, which is listed as $v_{\rm gal}$ in Table 1 of \citet{Free11}. $^c$~Upper~limit,~assuming~$v_{\rm jet} = c$ $^d$~For S2, velocity is based on the difference in redshift between the source galaxy and the group, not the velocity dispersion. \end{minipage} \end{table*} \subsection{Effects of Observational Resolution} Although the jets are well resolved in our simulations, they are typically unresolved in radio observations. From eqn.~\ref{eqn:r_over_h} and eqn.~\ref{eqn:pmin}, the density derived from observations will scale with jet thickness and radius of curvature as $n_{\rm IGM} \propto (h/R)^{-1/7} R^{-8/7}$. The values of $h/R$ in our simulations (table~\ref{table:model_data}) range from about $1/13$ to $1/21$ with a typical value of about $1/18$. The observed ratio (table~\ref{table:sources_emily}) ranges from $h/R = 1/14$ for S3 to $h/R = 1/0.83$ for S7. Although the density scales weakly with jet thickness, if we assume a real value of $h/R = 1/17$, corresponding to an initial jet opening angle of $\theta = 20\degr$, the densities for observed sources would be between 3\% (S3) and 54\% (S7) higher. If a $\theta = 5\degr$ inital opening angle is assumed, the correction would be between 52\% (S3) and 128\% (S7) For an under-resolved jet, the observed thickness will always be too high, and therefore the density derived will always be lower than the actual value. This is generally a fairly small correction, about 25\% for a typical source, but can be up to 50\% or more for sources with $h_{\rm obs} \sim R$. Inadequate resolution can also affect the measurement of the radius of curvature. To characterize this, we apply a Gaussian smoothing filter to our radio images of simulation .25E to simulate radio beam FWHM sizes from 1~kpc to 32~kpc, and then used our fitting routine to find the radius of curvature. The results are plotted in Fig.~\ref{fig:curve_vs_beam_size}. For this simulation, the actual radius of curvature is $R=36$~kpc and the jet thickness is $h=2$~kpc. For a marginally resolved jet (beam size $\leq 2$~kpc) the measured value of $R$ does not change. For larger beam sizes, $R$ is over-estimated by about 10\% to 25\%. As the beam size becomes comparable to the radius of curvature, the fit for $R$ becomes very poor (error comparable to $R$). Even in this case, however, the measured value of $R$ is systematically larger than the real value. An over-estimate of $R$ will lead to an underestimate of the density derived from eqn.~\ref{eqn:r_over_h}, typically about 20\% for sources with unresolved jets. Combining the effects of over-estimating the jet thickness and over-estimating the radius of curvature, we find that for a typical source in \citet{Free11} (resolved source, unresolved jet, $h/r \approx 1/4$) the density calculated is low by about 50\%, assuming a true value of $h/R$ of $1/17$. This ranges from no correction for source S3 (marginally resolved jet) to about 85\% low for source S7 (beam size $\sim R$). For estimates of the jet luminosity, the error due to resolution can be significantly larger. From eqn.~\ref{eqn:L_fit_pres}, we find that $L_{\rm jet} \propto h^{2} P_{\rm jet} \propto h^{6/7}$. There is no dependance on $R$, but luminosity is very sensitive to $h$. For example, assuming a real value of $h/R$ of $1/17$, the derived luminosity in table~\ref{table:sources_emily} would be reduced by a factor ranging from 1.2 (S3) to 13 (S7). \subsection{Effect of Viewing Angles \label{sect:viewing_angle}} So far, we have produced synthetic images of our simulations assuming that both the direction of jet propagation and the direction of motion of the AGN relative to the IGM are perpendicular to the observer's line of sight. In reality, for observed bent-double radio sources there will be unknown angles between the jet direction and the observer and the direction of motion and the observer, both of which will affect the measurement of the radius of curvature. The angle between the jet and the direction of motion will change where the point of maximum curvature is along the jet, but will not change the radius of curvature at that point \citep[e.g.][]{Begel79}. To quantify how much error viewing angle introduces, we take the data from simulation .25E, rotate it by various angles, then use it to create synthetic radio images. We then measure the radius of curvature in each of these images by the procedure in sect.~\ref{sect:measure_r}. Fig.~\ref{fig:curve_vs_angle_jet} plots the measured radius of curvature for an angle between the jet direction and the observer between $0\degr$ and $90\degr$. In all cases the direction of motion of the AGN is perpendicular to the observer's line of sight. The solid line in Fig.~\ref{fig:curve_vs_angle_jet} is $R = R_{\rm (\theta=0)} \times \cos(\theta)$, which is the expected apparent radius of curvature of a simple rotated circle. The measured value of $R$ follows this line fairly well out to about $60\degr$, where the apparent radius is half the real radius. Beyond this, the measured radius does not get any smaller. This is because the radio lobes at the ends of the jets begin to overlap from the observer's point of view, and their size dominates the fitting routine. This is one possible explanation for the appearance of source S7 in \citet{Free11}, which appears to have a small radius of curvature but with very thick jets. At $90\degr$, the fitting routine fails. However, it is unlikely that a source with the jet aimed almost directly at the observer would be classified as a bent-double radio source. Fig.~\ref{fig:curve_vs_angle_motion} plots the measured radius of curvature for an angle between the direction motion of the AGN and the observer between $0\degr$ and $90\degr$. In all cases the jet direction is perpendicular to the observer. The solid line is Fig.~\ref{fig:curve_vs_angle_motion} is $R = R_{\rm (\phi=0)} / \cos(\phi)$, which is the expected apparent radius of curvature of a simple circle rotated in the same manner. The measured value of $R$ follows this line out to about $60\degr$, where the apparent radius is double the real radius. Beyond this, the source is so inclined that the jet appears the be approximately straight and the fitting routine fails. In practice, sources will be rotated around both axes, and orientation effects tend to cancel each other out somewhat. Assuming sources rotated beyond $60\degr$ in either direction will not be classified as bent-double radio sources, the unknown viewing angle still introduces a large uncertainty in the measurement of $R$, and therefore in the values of $n_{\rm IGM}$ and $L_{\rm jet}$. On average, the error in the measurement of $R$ due to viewing angle should be between 50\% low to 30\% high ($1$-$\sigma$), but for an individual source could be up to a factor of 2 (100\% under-estimated to 50\% over-estimated). \subsection{X-ray Detectability} In general, the gas in galaxy groups is too cool and too low density to be seen in X-ray observations. However, as an AGN moves through the IGM, a cocoon of shocked material develops around the jets. The shock heats and compresses the IGM, boosting the X-ray emission. The parameters that affect the X-ray brightness are the velocity of the AGN, $v_{\rm gal}$, which determines the temperature of the shocked gas, and the density of the IGM, $n_{\rm IGM}$, which will determine the emissivity at a given temperature. To determine if bent-double radio sources in groups would be detectable in X-rays, we used the XIM tool \citep{Heinz09} to model 100~ks {\it Chandra} observations of four of our simulations, models .015E\_.25vel, .062E\_.5vel, .25E, and 1E\_4n, placed at a redshift of $z=0.1$. Fig.~\ref{fig:xray_images_all} shows synthetic {\it Chandra} images. In all figures we assume a background IGM temperature of $5\times10^5$~K and an IGM metallicity of $Z = 0.3$ solar. All four simulations have about the same radius of curvature of $R\approx36$~kpc. The first 3 simulations have the same IGM density with AGN velocities of $v_{\rm gal} = 250$~km/s (upper left), $500$~km/s (upper right) and $1000$~km/s (lower left). The last two simulations have the same AGN velocity ($v_{\rm gal} = 1000$~km/s) but different densities of $n_{\rm IGM} = 10^{-3}$~cm$^{-3}$ (lower left) and $4\times10^{-3}$~cm$^{-3}$ (lower right). For these images, we assume there is no point-source emission from the AGN. Even if there is point-source emission, however, the high spatial resolution provided by {\it Chandra} would be able to remove this contribution. Except in the lowest velocity case, there is significant X-ray emission from the region of the AGN jet and extended tail. We would expect emission to be strongest at the leading edge of the jet, where we are seeing the strongest part of the shock edge-on. However, because the edge-on shock is very thin and significant smoothing is needed to bring out the X-ray emission ($10$~arcsec in these images), the leading edge does not stand out. Instead, X-ray emission is spread across the region affected by the AGN and dominated by face-on shocks. In Fig.~\ref{fig:xray_plots_all}, we plot the mean surface brightness between $-50$ and $+50$ arcsec of the AGN along the y-axis (parallel to the jet) in Fig.~\ref{fig:xray_images_all}. In the worst-case presented here (model .015E\_.25vel, upper left panel), the surface brightness is about $0.02$~counts/arcsec$^2$ for a $100$~ks exposure, about twice the background level. As the AGN velocity increases, the surface brightness increases as roughly $v_{\rm gal}^{\sim0.75}$, reaching about $0.045$ counts/arcsec$^2$ for model .25E. For model 1E\_4n (lower right panel), which has the same velocity as model .25E, the brightness increases to $0.3$ counts/arcsec$^2$, a scaling of about $n_{\rm IGM}^{\sim1.5}$. In all cases, the shocks surrounding the AGN jet and tails are well resolved, so numerical mixing should have a minimal impact on the derived X-ray brightness. If these models were placed at a higher redshift, the angular size of the X-ray source would decrease, but the surface brightness would remain the same. For a large radius of curvature, the thickness of the shocked material, and therefore the surface brightness, will increase proportional to R. X-ray observations of bent-double radio sources in groups would place important constraints on the IGM properties. The X-ray surface brightness, along with an estimate of $v_{\rm gal}$ from velocity dispersion in the group, would allow $n_{\rm IGM}$ to be calculated independent of the radio observations. Density calculated this way would also be less sensitive to the AGN velocity, scaling as about $n_{\rm IGM} \propto v_{\rm gal}^{\sim-0.5}$ rather than $n_{\rm IGM} \propto v_{\rm gal}^{-2}$ as in eqn.~\ref{eqn:r_over_h}. The different scaling also means that in cases where $v_{\rm gal}$ is unconstrained, X-ray and radio data can be combined to find both the IGM density and AGN velocity. This can also be used to refine the vales of $v_{\rm gal}$ and $n_{\rm IGM}$ and constrain the viewing angle in cases where there is an estimate for the AGN velocity. Because the X-ray emission follows the tail of extended radio emission, measuring the length of the X-ray emitting region would provide an age estimate for the source. The general formula for the X-ray surface brightness is \begin{eqnarray} S \simeq 0.1 \times (Z+0.1) \times v_{\rm gal,1000}^{0.75} \times n_{\rm -3}^{1.5} \times ( R/36 \textrm{~kpc} ) \nonumber \\ \textrm{~counts/arcsec}^2 \label{eqn:xray_brightness} \end{eqnarray} \noindent for a 100~ks observation, where $Z$ is the metallicity of the IGM relative to solar. For sources S1 and S2 in table \ref{table:sources_emily} there are {\it Chandra} observations \citep{Free08} which found a total of $80\pm35$ and $94\pm26$ counts above the background in a 35.17~ks and 47.19~ks exposure, respectively. Assuming the total area on the sky of these sources is about $4R^2$ (the actual size of the emitting region is unknown), the total counts predicted would be very roughly 30 counts each for $Z=0.3$. This is lower than the observed counts, but current observations are not good enough to distinguish counts from the region of the bent-double from rest of the IGM, so they are still consistent. For particularly bright sources, there may be enough photons to determine the temperature of the shock. The temperature will scale roughly as $T \propto v_{\rm gal}^2$, so if the temperature can be found it would allow $v_{\rm gal}$ and $n_{\rm IGM}$ to be estimated from the X-ray data alone. With additional constraints from radio and velocity dispersion observations, it would then be possible to find the compression ratio of the shock, and from that the temperature of the unshocked intra-group medium. \section{Conclusions} Our simulations are able to closely match the appearance of observed bent-double radio sources with narrow, curved jets ending in bright radio lobes. We also predict that there should be an extended tail of radio emission that may be observable at low resolution and/or low frequencies. The length of this tail would allow the age of the AGN to be determined. From our simulations, we derive a formula for the radius of curvature (eqn.~\ref{eqn:radius_fit}) in terms of IGM density, AGN velocity, jet luminosity and jet velocity. From this we are able to calculate the kinetic jet luminosity (eqn.~\ref{eqn:L_fit_pres}) for observed bent-double radio sources, and find that $L_{\rm jet}$ is typically around $10^{45}$~erg/s, assuming that $v_{\rm jet} \simeq c$ and that the jets are really as thick as their observed values (see table~\ref{table:sources_emily}). The luminosities should be considered upper limits, as they will be lower if either the jets are slower or the jets are intrinsically thinner. This formula is independent of $v_{\rm gal}$ and therefore can be used to find $L_{\rm jet}$ even when the velocity of the AGN is unknown. A lack of resolution in radio observations leads to a systematic under-estimate of the IGM density for two reasons: the jet is unresolved and the radius of curvature is over-estimated. In our simulations, use initial conditions that produce ratio for jet thickness to radius of curvature of $h/R \simeq 1/17$, equivalent to a $20\degr$ initial opening angle of the jet. For observed sources (limited by resolution), this ratio ranged from $1/14$ to $1/0.83$ in \citet{Free11}. Density scales as $(h/R)^{-1/7}$, so this leads to an under-estimate of the IGM density of up to $50\%$. Inadequate resolution also leads to an over-estimate of the radius of curvature (and corresponding under-estimate of density), probably by about $20\%$. Overall, IGM density estimates for typical sources in \citet{Free11} are low by about $50\%$ due to resolution effects. This ranges from no correction for source S3 (assuming a marginally resolved jet) to about 85\% low for source S7 (beam size $\sim R$). The largest source of error, however, comes from the unknown angles between the observer, the jet direction and direction of motion of the AGN. This can lead to either an over- or under-estimate of the true radius of curvature. This leads to an uncertainty of up to a factor of 2 in the estimate of the IGM density, with a typical ($1$-$\sigma$) error or about $50\%$. This is comparable to the error from all observational uncertainties (dominated by the uncertainty in the AGN velocity). Even this uncertainty, however, does not change the conclusion that bending can be used to diagnose $\rho_{\rm IGM}$ in a statistical sense. Finally, we have modeled the X-ray emission of bent-double radio sources and predict that they should be detectable in {\it Chandra} X-ray observations. Although the IGM in groups is generally too cool and diffuse to be seen in X-ray observations, the shocks around the jet and tail in our simulations compress and heat the IGM, potentially making the entire region affected by the AGN jets detectable in X-rays. The X-ray surface brightness scales as approximately $S \propto n_{\rm IGM}^{1.5} v_{\rm gal}^{0.75}$. Sources in fairly dense environments and with fairly large angular size should be detectable in moderate ($\sim100$~ks) observations. Count rates from existing short X-ray observations of sources S1 and S2 in \citet{Free08} are consistent with our predictions. Further X-ray observations would provide complimentary constraints on the IGM density and AGN velocity to radio observations. For particularly bright sources, it may be possible to obtain a measure of the temperature of the shock from X-ray observations. This would provide an independent measure of the AGN velocity relative to the IGM. Future X-ray and radio observations will be able to place better constraints on the IGM density, AGN velocity and AGN age. With very complete observations it should also be possible to constrain the temperature of the IGM and the orientation of the jets and direction of AGN motion. \section*{Acknowledgments} BJM is supported by an NSF Astronomy and Astrophysics Postdoctoral Fellowship under award AST1102796. BJM and SH acknowledge NSF grant AST0707682. SH acknowledges NSF grant AST1109347. MB acknowledges support by the research group FOR 1254 funded by the Deutsche Forschungsgemeinschaft (DFG). MR acknowledges NSF grant 1008454. The software used in this work was in part developed by the DOE NNSA-ASC OASCR Flash Center at the University of Chicago. Resources supporting this work were provided by the NASA High-End Computing (HEC) Program through the NASA Advanced Supercomputing (NAS) Division at Ames Research Center. \bibliographystyle{mn2e}
\chapter{} \label{Appendix} \section{Discrete harmonic functions} The purpose of this chapter is to present the proofs of the maximum principle and the Harnack inequality (Theorem \ref{ERharnack}) for discrete harmonic functions. The Harnack inequality was used in Section \ref{ERsecreg} (in the proof of Lemma \ref{ERl4}) to construct the regeneration times. These inequalities are due to Kuo and Trudinger \cite{KT}. For the purpose of self-containedness, we will give the complete proofs of these estimates. We follow the arguments in \cite{KT}, adding to it some extra details. Recall the definitions of $a$, $L_a$, $b$ and $b_0$ in Section \ref{SeTriid1}. We consider discrete difference operates $L_a$ such that \[\sum_y a(x,y)=1, \quad\forall x,\] and $a(x,y)> 0$ only if $|x-y|=1$, denoted $x\sim y$. We assume that $L_a$ is uniformly elliptic with constant $\kappa\in(0,\frac{1}{2d}]$, that is, \[ a(x,y)\ge \kappa \text{ for any $x, y$ such that $x\sim y$}. \] For $r>0, x\in\mathbb{R}^d$, let $B_r(x)=\{z\in\mathbb{Z}^d: |z-x|<r\}$. We also write $B_r(o)$ as $B_r$. \subsection{Maximum principle} For any bounded set $E\subset\mathbb{Z}^d$, let $\partial E=\{y\in E^c:x\sim y \text{ for some }x\in E\}$, $\bar{E}=E\bigcup\partial E$ and $\mathop{\rm diam} E=\max\{|x-y|_\infty: x,y\in E\}$. \begin{theorem}\cite[Theorem 2.1]{KT}\label{AMP} Let $E\subset\mathbb{Z}^d$ be bounded and $u$ be a function on $\bar{E}$. For $x\in E$, define \[ I_u(x)=\{s\in\mathbb{R}^d: u(x)-s\cdot x\ge u(z)-s\cdot z, \forall z\in\bar{E}\}. \] If \[L_a u(x)\ge -g(x)\] for all $x\in E$ such that $I_u(x)=I_u(x,E,a)\neq \emptyset$, then \[ \max_E u\le C\mathop{\rm diam}(\bar{E}) \big(\sum_{x\in E, I_u(x)\neq \emptyset}|g|^d\big)^{1/d}+\max_{\partial E}u, \] where $C$ is a constant determined by $d, \kappa$ and $b_0\mathop{\rm diam} E$. \end{theorem} {\it Proof:}~ Without loss of generality, we assume $g\ge 0$ and \[ \max_E u=u(x_0)>\max_{\partial E}u \] for some $x_0\in E$. Otherwise, there is nothing to prove. For $s\in \mathbb{R}^d$ such that \begin{align}\label{A*1} |s|_\infty &\le[u(x_0)-\max_{\partial E}u]/(d\mathop{\rm diam}\bar{E})\nonumber\\ &=:R=R(u,E), \end{align} we have \[ u(x_0)-u(x)\ge s\cdot(x_0-x) \] for all $x\in\partial E$, which implies that $\max_{z\in\bar{E}}u(z)-s\cdot z$ is achieved in $E$. Hence $s\in \bigcup_{x\in E}I_u(x)$ and the cube \[ Q_R:=\{x:|x|_{\infty}< R\}\subset \bigcup_{x\in E}I_u(x). \] For any $p\in\mathbb{R}^d$, set \[ f(p)=(|p|^{d/d-1}+\mu^{d/d-1})^{1-d}, \] where $\mu>0$ is a constant to be fixed later. Since for any $x\in E$, $I_u(x)\subset\mathbb{R}^d$ is bounded and closed, we can choose $p_x\in I_u(x)$ so that \[ |p_x|=\min_{p\in I_u(x)}|p|. \] Then \[ f(p_x)=\max_{p\in I_u(x)}f(p). \] Thus \begin{equation}\label{A*2} \int_{Q_R}f(s)\, \mathrm{d} s\le \int_{\bigcup_{x\in E}I_u(x)}f(s)\, \mathrm{d} s \le \sum_{x:I_u(x)\neq\emptyset}f(p_x)|I_u(x)|, \end{equation} where $|I_u(x)|$ denotes the Lebesgue measure of $I_u(x)$. Further, we will show that, for any $x\in E$ with $I_u(x)\neq\emptyset$, \begin{equation}\label{A*14} |I_u(x)|\le (2/\kappa)^{d}[g(x)+b(x)p_x]^d. \end{equation} To this end, we fix an $x\in E$ with $I_u(x)\neq\emptyset$ and set \[ w(z)=u(z)-p_x(z-x), \quad\forall z\in\bar{E}. \] Then $w(x)\ge w(z)$ for all $z\in\bar{E}$ and \begin{equation}\label{A*3} I_u(x)=I_w(x)+p_x. \end{equation} Since for any $q\in I_w(x)$ and $i=1,\ldots,d$, \[ w(x)-w(x\pm e_i)\ge\mp q_i, \] we obtain (by ellipticity and by $w(x)\ge w(z)$, $\forall z\in\bar{E}$) \begin{align*} 0\le \kappa|q|_\infty &\le\sum_{y}a(x,y)(w(x)-w(y))\\ &=-L_a u+b(x)p_x\\ &\le g(x)+b(x)p_x. \end{align*} Hence \[ I_w(x)\subset \big[\frac{-g(x)-b(x)p_x}{\kappa},\frac{g(x)+b(x)p_x}{\kappa}\big]^d \] and \[ |I_u(x)|\stackrel{(\ref{A*3})}{=}|I_w(x)|\le (2/\kappa)^{d}[g(x)+b(x)p_x]^d. \] (\ref{A*14}) is proved. (\ref{A*14}) and (\ref{A*2}) yield \[ \int_{Q_R}f(s)\, \mathrm{d} s \le (\dfrac{2}{\kappa})^d\sum_{x:I_u(x)\neq\emptyset}f(p_x)[g(x)+b(x)p_x]^d. \] Since by H\"{o}lder's inequality, \[ g(x)+|b(x)||p_x| \le \big[(\frac{g(x)}{\mu})^d+|b(x)|^d\big]^{1/d} \big[\mu^{d/d-1}+|p_x|^{d/d-1}\big]^{(d-1)/d}, \] we get \begin{equation}\label{A*4} \int_{Q_R}f(s)\, \mathrm{d} s \le (\frac{2}{\kappa})^d\sum_{x:I_u(x)\neq\emptyset} \big[(\frac{g(x)}{\mu})^d+|b(x)|^d\big]. \end{equation} On the other hand, by H\"{o}lder's inequality, \[ f(s)=(|s|^{d/d-1}+\mu^{d/d-1})^{1-d}\ge 2^{2-d}(|s|^d+\mu^d)^{-1}. \] Thus \begin{align}\label{A*5} \int_{Q_R}f(s)\, \mathrm{d} s \ge \int_{B_R}f(s)\, \mathrm{d} s &\ge 2^{2-d}\int_{B_R}(|s|^d+\mu^d)^{-1}\, \mathrm{d} s\nonumber\\ &=2^{2-d}\frac{\mathcal{O}_d}{d}\log[(\frac{R}{\mu})^d+1], \end{align} where $\mathcal{O}_d$ is the area of the unit sphere in $\mathbb{R}^d$. Finally, combining (\ref{A*4}) and (\ref{A*5}) and putting \[ \mu:=[\sum_{x:I_u(x)\neq\emptyset}g(x)^d]^{1/d},\] we conclude that \[ \kappa^d 2^{2-2d}\frac{\mathcal{O}_d}{d}\log[(\frac{R}{\mu})^d+1] \le 1+(b_0\mathop{\rm diam}\bar{E})^d. \] Recalling the definition of $R=R(u,E)$ in (\ref{A*1}), the theorem follows. \qed By the same argument as in the proof of Theorem \ref{Cmvi} (Section \ref{SeTriid1}), Theorem~\ref{AMP} and Lemma~\ref{Cmvilemma} imply \begin{theorem}[Mean-value inequality]\label{Amvi} For any function $u$ on $\bar{B}_R$ such that \[ L_a u\ge 0, \quad x\in B_R \] and any $\sigma\in (0,1)$, $0<p\le d$, we have \[ \max_{B_{\sigma R}}u\le C\norm{u^+}_{B_R,p}, \] where $C$ depends on $\sigma, p, \kappa, d$ and $b_0R$. \end{theorem} \subsection{Harnack inequality} \begin{theorem}[Harnack inequality]\cite[Corollary 4.5]{KT}\label{ERharnack} Let $u$ be a non-negative function on $B_R$, $R>1$. If \[ L_a u=0 \] in $B_R$, then for any $\sigma\in (0,1)$ with $R(1-\sigma)>1$, we have \[ \max_{B_{\sigma R}}u\le C\min_{B_{\sigma R}}u, \] where $C$ is a positive constant depending on $d, \kappa, \sigma$ and $b_0 R$. \end{theorem} \begin{lemma}\label{Ahlemma} Suppose $u$ is a non-negative function on $\bar{B}_R$ that satisfies \[ L_a u\le 0\] in $B_R$. Then for any $\sigma\le\tau<1$, \begin{equation}\label{Aehlemma} \min_{B_{\tau R}}u\ge C\min_{B_{\sigma R}}u, \end{equation} where $C$ depends on $\kappa,d, \sigma,\tau$ and $b_0R$. \end{lemma} {\it Proof:}~ Recall the definition of $\eta=\eta_R(x)$ in Lemma \ref{Cmvilemma}. We will first show that there exists a constant $\beta=\beta(\sigma, b_0R, \kappa)$ such that \begin{equation}\label{A*10} L_a \eta\ge -(2^\beta+\beta^3) R^{-3} \qquad\text{ in }B_R\setminus B_{\sigma R}. \end{equation} If $R-1\le |x|< R$, then $\eta(x)\le (2/R)^\beta\le 2^\beta R^{-3}$ for $\beta\ge 3$. Hence for $\beta\ge 3$, \[L_a\eta\ge -\eta\ge -2^\beta R^{-3}.\] If $\sigma R\le |x|<R-1$, then $y\in B_R$ for all $y\sim x$. For $i=1,\ldots, d$, the third derivative $D_i^3\eta$ of $\eta$ with respect to $x_i$ satisfies \begin{align*} |D_i^3\eta| &=\big|4\beta(\beta-1)x_iR^{-4}\eta^{1-3/\beta} [3(1-|x|^2 /R^{2})-2(\beta-2)x_i^2/R^2]\big|\\ &\le 4\beta(\beta-1)(2\beta-1)R^{-3}, \end{align*} and so, by Taylor's expansion, \[ \eta(x+e)-\eta(x)\ge \nabla\eta(x)\cdot e+\frac{1}{2}e^{T}D^2\eta(x)e-\frac{8}{6}\beta^3 R^{-3}. \] Thus \begin{align*} L_a\eta(x) &=\sum_{e}a(x,e)(\eta(x+e)-\eta(x))\\ &\ge \nabla\eta\cdot b(x)+\frac{1}{2}\sum_e a(x,x+e)e^{T}D^2\eta(x)e-\frac{4}{3}\beta^3 R^{-3}. \end{align*} Noting that \[ \nabla\eta\cdot b(x) \stackrel{\eqref{Afifth}, \eta\le 1}{\ge} -2(b_0R)\beta R^{-2}\eta^{1-2/\beta}, \] and, for $\sigma R\le |x|<R-1$, \begin{align*} &\sum_e a(x,x+e)e^{T}D^2\eta(x)e\\ &=\sum_{i=1}^d(a(x,x+e_i)+a(x,x-e_i))D_{ii}\eta(x)\\ &=2\beta R^{-2}\eta^{1-2/\beta}\sum_{i=1}^d\big(a(x,x+e_i)+a(x,x-e_i)\big)\big(\frac{2(\beta-1)x_i^2}{R^2}-(1-\frac{|x|^2}{R^2})\big)\\ &\ge 2\beta R^{-2}\eta^{1-2/\beta}[4\kappa(\beta-1)\sigma^2-1], \end{align*} we have \[ L_a\eta \ge [4\kappa(\beta-1)\sigma^2-1-2b_0R]R^{-2}\eta^{1-2/\beta} -\frac{4}{3}\beta^3R^{-3}. \] Hence \eqref{A*10} also holds for $\sigma R\le |x|<R-1$ if we take \[ \beta\ge 1+\frac{1+2b_0R}{4\kappa\sigma^2}. \] \eqref{A*10} is proved. Next, let $m_\sigma:=\min_{B_{\sigma R}}u$ and $w:=m_\sigma\eta-u$. Then \begin{equation}\label{A*11} \max_{B_{\tau R}}w\ge (1-\tau^2)^\beta m_\sigma-m_\tau. \end{equation} Since $w\le 0$ in $B_{\sigma R}\bigcup B_R^c$ and \[ L_a w\stackrel{\eqref{A*10}}{\ge} -(2^\beta+\beta^3)m_\sigma R^{-3} \qquad\text{in }B_R/B_{\sigma R}, \] we get by the maximum principle that \begin{equation}\label{A*12} \max_{B_R}w \le C_1 m_\sigma R^{-1}, \end{equation} where $C_1$ depends on $\kappa,d,\sigma$ and $b_0R$. By \eqref{A*11} and \eqref{A*12}, \[ [(1-\tau^2)^\beta-\frac{C_1}{R}]m_\sigma\le m_\tau. \] Therefore, \eqref{Aehlemma} holds if $R$ satisfies \[R>\frac{2C_1}{(1-\tau^2)^\beta}.\] For $R\le\frac{2C_1}{(1-\tau^2)^\beta}$, it follows by iteration (noting $\kappa u(x)\le u(y)$ for $x\sim y$) that \[ \kappa^{2C_1(1-\tau^2)^{-\beta}}m_\sigma\le m_\tau. \] \eqref{Aehlemma} is proved.\qed\\ For any $z\in\mathbb{Z}^d$ and any $n=(n_1,\ldots,n_d)\in\mathbb{N}^d$ , we let \[ N(z,n):=(z+\prod_{i=1}^d[0,n_i-1])\cap\mathbb{Z}^d. \] We say that $N(z,n)$ is \textit{nice} if $n$ satisfies $\max_{i,j}|n_i-n_j|\le 1$. Call $|n|_\infty$ the \textit{length} of the nice rectangle $N(z,n)$. Intuitively, a nice rectangle is ``nearly a cube". \begin{proposition}\label{Aprop0} Let $u$ be a nonnegative function on $\bar{B}_R, R>0$ such that \[ L_a u\le 0 \quad\text{ in }B_R. \] Suppose $r\in(0, R/7\sqrt{d}]$ and $N=N(z,n)\subset Q_r$ is a nice rectangle in $Q_r$. Then there exists a constant $\delta=\delta(d,\kappa,b_0R)\in(0,1)$ such that, if $\Gamma\subset B_R$ satisfies \[|\Gamma\cap N|\ge \delta|N|,\] then \[ \min_{N'}u\ge C\min_{\Gamma}u, \] where $N'=(z+\prod_{i=1}^d[-n_i,2n_i-1])\cap\mathbb{Z}^d$ and $C$ depends on $\kappa,d, \sigma, \tau$ and $b_0R$. \end{proposition} {\it Proof:}~ When $|n|_\infty=1$, $N$ is a singleton, and the proposition follows by iteration (noting that $u(x)\le\kappa u(y)$ for any $x\sim y$). So we only consider the case when the length of $N$ is $\ge 2$. \begin{figure} \centering \includegraphics[width=0.6\textwidth]{boxes} \caption{$N$ is the rectangle in the center. $B_{h}(O_N)$ is the small circle.} \label{Afig2} \end{figure} Denote the center of $N$ by $O_N=z+(\frac{n_1-1}{2},\cdots,\frac{n_d-1}{2})\in(\frac{1}{2}\mathbb{Z})^d$. Setting $h=:\min_{i}n_i/2$, we have \[ B_h(O_N)\subset N\subset B_{2\sqrt{d}h}(O_N). \] Since $h\ge\frac{|n|_\infty-1}{2}\ge\frac{|n|_\infty}{4}$, we have \[ N'\subset B_{3|n|_\infty\sqrt{d}/2}(O_N) \subset B_{6\sqrt{d}h}(O_N). \] Suppose for some $\delta\in(0,1)$, \[|\Gamma\cap N|\ge \delta|N|.\] Let $u_\Gamma=:\min_\Gamma u$ and \[v=:u_\Gamma-u,\] then $L_a v\le 0$ and $v^+|_\Gamma=0$. By Theorem \ref{Amvi}, \begin{align*} \max_{B_{h/2}(O_N)}v &\le C\frac{1}{|B_{h}|}\sum_{B_h(O_N)}v^+\\ &\le C\frac{|B_h(O_N)\setminus\Gamma|}{|B_h|}\max_{B_h(O_N)}v\\ &\stackrel{|B_h|\ge C|N|}{\le} C\frac{|N\setminus\Gamma|}{|N|}\max_{B_h(O_N)}v \le C_2(1-\delta)\max_{B_h(O_N)}v, \end{align*} where $C_2$ depends on $\kappa,d$ and $b_0R$. Taking $\delta=\delta(\kappa, d,b_0R)$ big enough such that $C_2(1-\delta)\le 1/2$, we get \[\max_{B_{h/2}(O_N)}v \le \frac{1}{2}\max_{B_h(O_N)}v.\] Hence \[ u_\Gamma-\min_{B_{h/2}(O_N)}u\le\frac{1}{2}(u_\Gamma-\min_{B_h(O_N)}u). \] Therefore, noting that (since $r\le R/7\sqrt{d}$) $B_{7\sqrt{d}h}(O_N)\subset B_R$, \[ u_\Gamma\le 2\min_{B_{h/2}(O_N)}u \stackrel{\text{Lemma \ref{Ahlemma}}}{\le} C \min_{B_{6\sqrt{d}h}(O_N)}u\le C \min_{N'}u, \] with $C$ depending on $\kappa,d$ and $b_0R$.\qed\\ \begin{lemma}\label{Ahlemma2} Let $u$ be a nonnegative function on $\bar{B}_R, R>0$ such that \[ L_a u\le 0 \quad\text{ in }B_R. \] Let $r\in(0, R/7\sqrt{d}]$. Then for any $\Gamma\subset Q_r$, there exists a subset $\Gamma_\delta\supset\Gamma$ of $Q_r$ such that either $\Gamma_\delta=Q_r$ or $|\Gamma_\delta|>\delta^{-1}|\Gamma|$ holds, and \[ \min_{\Gamma_\delta}u\ge\gamma\min_\Gamma u. \] Here the constant $\gamma$ depends only on $\kappa, d$ and $b_0R$, and $\delta$ is the same as in Proposition \ref{Aprop0}. \end{lemma} {\it Proof:}~ We will construct $\Gamma_\delta$ through a cube decomposition procedure. Observe that any nice rectangle with length $l\ge 2$ can be decomposed into (at most $2^d$) smaller disjoint nice rectangles whose lengths are either $\lfloor \frac{l}{2}\rfloor$ or $\lfloor \frac{l}{2}\rfloor+1$. With abuse of terminology, we say that such a decomposition is \textit{nice}. Note that a nice decomposition may not be unique. For any $\Gamma\subset Q_r$, set \[\mathcal{N}=\mathcal{N}(\Gamma):=\{N:N \text{ is nice and }|\Gamma\cap N|\ge \delta |N|\}.\] Now perform cube decompositions to $Q_r$ as follows. Assume that we have an imaginary ``bag". In the first step, we put $Q_r$ into our ``bag" if $Q_r\in\mathcal{N}$, and decompose $Q_r$ nicely (into at most $2^d$ nice rectangles) if otherwise. In the second step, we repeat the same procedure on each of the remaining rectangles, i.e., put a rectangle into our ``bag" if it is in $\mathcal{N}$, and decompose a rectangle (with lengths$\ge 2$) nicely if it is not in $\mathcal{N}$. Repeat this procedure as often as necessary, and stop if there is nothing to decompose or all the remaining rectangles are singletons in $Q_r\setminus\Gamma$. The process will end within finite number of steps. Denote the collection of the rectangles in our ``bag" by $\mathcal{N}_0$($\subset\mathcal{N}$). For $N\in\mathcal{N}_0$ and $N\neq Q_r$, we denote by $N^{-1}$ its \textit{prior}, i.e, $N$ is obtained from a nice decomposition of $N^{-1}$ in the previous step. Set $Q_r^{-1}=Q_r$ and \[ \Gamma_\delta:=\bigcup_{N\in\mathcal{N}_0}N^{-1}. \] Recall the definition of $N'$ in Proposition \ref{Aprop0}. For any $N\in\mathcal{N}_0$, since $|\Gamma\cap N|\ge \delta |N|$ and $N^{-1}\subset N'$, by the Proposition \ref{Aprop0} we have \[ \min_{N^{-1}}u \ge\min_{N'}u \ge\gamma\min_\Gamma u. \] Hence, \[ \min_{\Gamma_\delta}u\ge\gamma\min_\Gamma u. \] Moreover, note that $\Gamma_\delta=Q_r$ when $\mathcal{N}_0=\{Q_r\}$. Otherwise, if $\mathcal{N}_0\neq\{Q_r\}$, we have \[ |\Gamma\cap N^{-1}|<\delta|N^{-1}|\quad\text{for all } N\in\mathcal{N}_0. \] Therefore, if $\mathcal{N}_0\neq\{Q_r\}$, \begin{align*} |\Gamma|=\Abs{\bigcup_{N\in\mathcal{N}_0}(\Gamma\cap N)} &\le \Abs{\bigcup_{N\in\mathcal{N}_0}(\Gamma\cap N^{-1})}\\ &<\sum_{N^{-1}:N\in\mathcal{N}_0}\delta\abs{N^{-1}}=\delta\abs{\Gamma_\delta}. \end{align*} Our proof is complete. \qed\\ \noindent{\it Proof of Theorem \ref{ERharnack}:}\\ We only consider the case when $\sigma<1/7\sqrt{d}$. For any $\Gamma\subset Q_{\sigma R}$, if $|Q_{\sigma R}|\le\delta^{-s}|\Gamma|$ for some $s\in\mathbb{N}$, then we have \[ m:=\min_{Q_{\sigma R}}u\ge\gamma^s\min_\Gamma u \] by Lemma \ref{Ahlemma2} and iteration. Hence for $t\ge 0$, putting $\Gamma^t:=\{x\in Q_{\sigma R}: u(x)\ge t\}$, we get \begin{equation}\label{A*13} m\ge\gamma^{\lceil\log_\delta(|\Gamma^t|/|Q_{\sigma R}|)\rceil}t \ge \left(\frac{|\Gamma^t|}{|Q_{\sigma R}|}\right)^{\log_\delta\gamma}\gamma t. \end{equation} Note that $q:=\log_\gamma \delta>0$, since $\gamma,\delta\in(0,1)$. Therefore, for any $p\in(0,q)$, \begin{align*} \frac{1}{|Q_{\sigma R}|}\sum_{Q_{\sigma R}}u^p &=m^p+\frac{1}{|Q_{\sigma R}|}\sum_{Q_{\sigma R}}\int_m^\infty pt^{p-1}1_{u\ge t}\, \mathrm{d} t\\ &=m^p+\int_m^\infty pt^{p-1}\frac{|\Gamma^t|}{|Q_{\sigma R}|}\, \mathrm{d} t\\ &\stackrel{(\ref{A*13})}{\le} m^p+\int_m^\infty pt^{p-1}(\frac{m}{\gamma t})^q\, \mathrm{d} t\le C m^p, \end{align*} where $C$ depends on $\kappa, d$ and $b_0R$. Combining this and Theorem \ref{Amvi}, the Harnack inequality for $\sigma<1/7\sqrt{d}$ is proved. The case $\sigma\ge 1/7\sqrt{d}$ then follows by a chaining argument. \qed \chapter{Invariance Principle for Random Walks in Balanced Random Environment} \label{CLT chapter} This chapter is devoted to the proofs of Theorem~\ref{CLT1} and Theorem~\ref{CLT2}. The structure of this chapter is as follows. In Section \ref{SePeEn} we construct the ``periodized environments" as in \cite{Sz1, ZO}, and show that the proof of $Q\sim P$ can be reduced to the proof of the inequality (\ref{CPhi0}). Using the maximum principle (Theorem \ref{CMP}), we then prove (\ref{CPhi0}) in Section \ref{SeMP} under the assumptions of Theorem \ref{CLT1}(i). In Section \ref{SePercE}, which is devoted to the iid setup, we prove Theorem \ref{CLT2}(i) using percolation tools. Section \ref{SeTran} is devoted to the proof of the transience of the RWRE for $d\geq 3$, thus providing a proof of Theorem \ref{CLT1}(ii). In Section \ref{SeTriid}, we will show a modified maximum principle for balanced difference operators, and use it to prove Theorem \ref{CLT2}(ii). Throughout, $C$ denotes a generic positive constant that may depend on dimension only, and whose value may change from line to line. \section {The periodized environments}\label{SePeEn} As in \cite{Sz1, ZO}, the following periodic structure of the environment is introduced. Let $ \Delta_N (x_0)=\{x\in \mathbb{Z}^{d}: |x-x_0|_{\infty}\le N\} $ be the cube centered at $x_0$ of length $2N$. Let $\Delta_N=\Delta_N(o)$. For any $ x\in\mathbb{Z}^{d} $, set $$ \hat{x}:=x+(2N+1)\mathbb{Z}^{d}\in \mathbb{Z}^{d}/(2N+1)\mathbb{Z}^{d}. $$ For any fixed $ \omega \in \Omega $, we define $\omega^{N}$ by setting $ \omega^{N}(x)=\omega(x) $ for $ x \in \Delta_{N} $ and $\omega^N (y)=\omega^N (x)$ for $y \in \mathbb{Z}^{d} $ whenever $\hat{y}=\hat{x}$. Let $ \Omega^{N}=\{\omega^{N}: \omega\in\Omega\}$. Let $ \{X_{n,N}\} $ denote the random walk on $ \mathbb{Z}^{d} $ in the environment $ \omega^{N} $. Then $ \{\hat{X}_{n,N}\} $ is an irreducible finite-state Markov chain, hence it possesses a unique invariant probability measure, which can always be written in the form \[ \dfrac{1}{(2N+1)^d}\sum_{x\in\Delta_{N}}\Phi_{N}(x)\delta_{\hat{x}} . \] Here $\Phi_N$ is some function on $\Delta_N$ and $(2N+1)^{-d}\Phi_{N}(\cdot)$ sums to $1$, so that $\Phi_N$ can be interpreted as a density with respect to the uniform measure on $\Delta_N$. Define \[ Q_{N}=Q_{N,\omega}=\dfrac{1}{(2N+1)^{d}}\sum_{x\in\Delta_{N}}\Phi_N (x)\delta_{\theta^{x}\omega^{N}} \] as a probability measure on $\Omega^N$. Then, for any $x\in \Delta_N$, \begin{align*} \sum_{y\in\Delta_N} Q_N(\theta^y \omega^N)M(\theta^y \omega^N, \theta^x\omega^N) &=\sum_{y\in\Delta_N} \frac{\Phi_N (y)}{(2N+1)^d}\omega^N(y,x)\\ &= \frac{\Phi_N (x)}{(2N+1)^d}=Q_N (\theta^x \omega^N). \end{align*} This implies that $Q_N$ is the invariance probability measure (with respect to the kernel $M$) for the Markov chain $\{\bar{\omega}^{N}(n)\}$ on $\Omega^{N}$. We will show that $Q_N$ converges weakly to some measure $ Q $ with good properties. To do this, we first introduce a sequence of measures \[ P_{N}=P_{N,\omega}=\dfrac{1}{(2N+1)^{d}} \sum_{x\in\Delta_{N}}\delta_{\theta^{x}\omega^{N}}, \] which, by the multidimensional ergodic theorem (see Theorem (14.A8) in \cite{Ge} and also Theorem 1.7.5 in \cite{Kr}), converges weakly to $P$, $P$-a.s. Let $ \{\omega_{\gamma}^{N}\}_{\gamma=1}^{k} $ denote the set of distinct states in $ \{ \theta^{x}\omega^{N}\}_{x\in \Delta_{N}} $ and $ C_{N}(\gamma):=\{x\in \Delta_{N}: \theta^{x}\omega^{N}=\omega_{\gamma}^{N}\} $. Set, for any finite subset $E\subset\mathbb{Z}^d$, \[ \lVert f\rVert_{E,j}:=(|E|^{-1}\sum_{x \in E} |f(x)|^{j})^{\frac{1}{j}}. \] Since $ \, \mathrm{d} Q_N/\, \mathrm{d} P_N=\sum_{\gamma=1}^{k}\delta_{\omega_{\gamma}^{N}}|C_{N}(\gamma)|^{-1}\sum_{ x\in C_{N}(\gamma)}\Phi_{N}(x):=f_{N}$, we have that, for any measurable function $g$ on $\Omega$, \begin{align}\label{Q_n0} |Q_N g| &\le (\int f_{N}^{\alpha} \, \mathrm{d} P_N)^{\frac{1}{\alpha}}(\int |g|^{\alpha'} \, \mathrm{d} P_N)^{\frac{1}{\alpha'}} \nonumber\\ &\le \big( \frac{1}{|\Delta_N |}\sum_{\gamma =1}^{k}\sum_{x \in C_N (\gamma)}\Phi_N (x)^{\alpha} \big)^{\frac{1}{\alpha}}(\int |g|^{\alpha'} \, \mathrm{d} P_N)^{\frac{1}{\alpha'}} \nonumber\\ &= \lVert \Phi _N \rVert _{\Delta_N, \alpha}(P_{N}|g|^{\alpha'})^{\frac{1}{\alpha'}}, \end{align} where $\alpha'$ is the H\"older conjugate of $\alpha$, $1/\alpha+1/\alpha'=1$, and we used H\"older's inequality in the first and the second inequalities. Since $\Omega$ is compact with respect to the product topology, along some subsequence $N_k\to\infty$, $\{Q_{N_k}\}$ converges weakly to a limit, denoted $Q$. Assume for the moment that \begin{equation}\label{CPhi0} \varlimsup_{N\to\infty}\lVert \Phi _N \rVert _{\Delta_N, \alpha} \le C, \quad P\mbox{- }a.s. \end{equation} We then show that, for a.e. $\omega\in\Omega$, \begin{equation}\label{CLTf2} Q\ll P . \end{equation} Indeed, let $A\subset \Omega$ be measurable. Let $ \rho $ denote a metric on the Polish space $ \Omega $. For any closed subset $ F \subset A $, $ \delta >0 $, introduce the function $f(\omega)=[1-\rho (\omega, F)/\delta ]^+$ which is supported on $ F_{\delta}=\{\omega \in \Omega: \rho (\omega, F) <\delta\}$. Then, by (\ref{Q_n0}), (\ref{CPhi0}), \begin{equation*} Q F\le \varlimsup_{N\to\infty}Q_N f \le C (P f^{\alpha'})^{\frac{1}{\alpha'}}\le C (P F_{\delta})^{\frac{1}{\alpha'}} . \end{equation*} Letting $ \delta\downarrow 0 $, we get $ Q F \le C (P F)^{\frac{1}{\alpha'}}$. Taking supremums over all closed subset $ F \subset A $, one concludes that $Q A \le C\cdot (P A)^{\frac{1}{\alpha'}}$, which proves (\ref{CLTf2}). Once we have (\ref{CLTf2}), it is standard to check, using ellipticity, that $ \bar{\omega}(n) $ is ergodic with respect to $ Q $ and $Q\sim P$ (see \cite{Sz1, ZO}). (Thus, by the ergodic theorem, $Q$ is uniquely determined by $Qg=\mathop{\rm lim}_{n\to\infty} E\sum_{j=0}^{n-1}g(\bar{\omega}_j)/n$ for every bounded measurable $g$. Hence $Q$ is \textit{the} weak limit of $Q_N$.) Therefore, to prove the invariance principle it suffices to prove (\ref{CPhi0}). Sections \ref{SeMP} and Section \ref{SePercE} are devoted to the proof of (\ref{CPhi0}), under the assumptions of Theorems \ref{CLT1} and \ref{CLT2}. \section {Maximum principle and proof of Theorem \ref{CLT1}(i)}\label{SeMP} Throughout this section, we fix an $\omega\in \Omega$. For any bounded set $ E \subset \mathbb{Z}^{d} $, let $\partial E =\{y \in E^{c}: \exists x\in E, |x-y| _{\infty}=1\}$, $ \bar{E}=E \bigcup \partial E $ and $ \mathop{\rm diam}(E)=\max\{|x-y| _{\infty}: x, y\in E\} $. For any function $f$ defined on $ \bar{E}$ , let $ L_{\omega} $ denote the operator \begin{equation}\label{operator} (L_{\omega}f )(x)=\sum_{i=1}^{d}\omega(x, e_i)[f(x+e_i)+f(x-e_i)-2f(x)], \quad x\in E . \end{equation} The following discrete maximum principle is an adaption of Theorem 2.1 of \cite{KT}. \begin{theorem}[Maximum Principle]\label{CMP} Let $E\subset \mathbb{Z}^d$ be bounded, and let $u$ be a function on $ \bar{E} $. For all $x\in E$, assume $ \varepsilon(x)>0 $ and define $$ I_{u}(x):=\{s \in \mathbb{R}^{d}: u(x)-s\cdot x \ge u(z)-s\cdot z, \forall z \in\bar{E}\} .$$ If $ L_{\omega} u(x) \ge -g(x)$ for all $x \in E$ such that $I_u (x)\ne \emptyset$, then \begin{equation}\label{Cmpremark} \max_{E} u \le C\mathop{\rm diam}{\bar{E}} \bigg(\sum_{\substack{ x\in E\\ I_u (x)\ne \emptyset} }|\frac{g}{\varepsilon}|^d\bigg)^{\frac{1}{d}} +\max_{\partial E}u . \end{equation} In particular, \[ \max_{E} u \le C\mathop{\rm diam}{\bar{E}}\cdot |E|^{\frac{1}{d}} \lVert \frac{g}{\varepsilon}\rVert _{E,d}+\max_{\partial E}u .\] \end{theorem} {\it Proof:}~ See the proof of Theorem 2.1 in \cite{KT}.\qed\\ Define the stopping times $ \tau_0=0 $, $ \tau_1 =\tau :=\min \{j \ge 1: |X_{j,N}-X_{0,N}|_{\infty}> N\} $ and $ \tau_{j+1}=\min \{n>\tau_j : |X_{n,N} -X_{\tau_j, N}|_{\infty}> N\} $. \begin{lemma}\label{CLTE} Let $\omega^{N}$, $\{X_{n,N}\}$ be as in Section 1 and $\tau$ as defined above, then there exists a constant $c$ such that, for all $N$ large, $$E_{\theta^{x}\omega^{N}}^{o} (1-\frac{c}{N^2})^{\tau}\leq C <1 .$$ \end{lemma} {\it Proof:}~ Since P is balanced, $ X_{n,N} $ is a martingale and it follows from Doob's inequality that for any $K \ge 1 $, \begin{align*} P_{\theta^{x}\omega^{N}}^{o} \{\tau \le K\} & \le 2 \sum_{i=1}^{d} P_{\theta^{x}\omega^{N}}^{o}\{\sup_{n \le K} X_{n, N}(i) \ge N+1\}\\ & \le \frac{2}{N+1} \sum_{i=1}^{d}E_{\theta^{x}\omega^{N}}^{o} X_{K, N} (i)^{+} \le \frac{2d}{N+1} \sqrt{K}, \end{align*} where $ X_{n,N} (i) $ is the $i$-th coordinate of $ X_{n,N} $. Hence \begin{equation*} E_{\theta^{x}\omega^{N}}^{o} (1-\frac{c}{N^2})^{\tau} \le (1-\frac{c}{N^2})^{K} +\frac{2d}{N+1}\sqrt{K} . \end{equation*} Taking $c=16 d^2$ and $K=N^2/16d^2$, we get $E_{\theta^{x}\omega^{N}}^{o} (1-\frac{c}{N^2})^{\tau}\le e^{-1}+2^{-1} .$ \qed \begin{theorem}\label{CPhi} \begin{equation}\label{CPhi1} \lVert \Phi_N \varepsilon\rVert _{\Delta_N , \beta}\le C, \end{equation} where $\beta=d'=d/(d-1)$. \end{theorem} {\it Proof:}~ Let $c$ be the same constant as in the previous lemma. For any function $ h\ge 0 $ on $\Delta_N $, \begin{align*} &\lVert \Phi_N \cdot h\rVert _{\Delta_N , 1}\\ &= \frac{c}{N^2}\sum_{x \in \Delta_N}\frac{\Phi_N (x)}{|\Delta_N |} \sum_{m\ge 0}E_{\omega^N}^{x}\sum_{\tau_m\le\ j <\tau_{m+1}}(1-\frac{c}{N^2})^{j}h(\hat{X}_{j,N})\\ &\le \frac{c}{N^2}\sum_{x \in \Delta_N}\frac{\Phi_N (x)}{|\Delta_N |}\sum_{m\ge 0}E_{\omega^N}^{x}(1-\frac{c}{N^2})^{\tau_m} E_{\omega^N}^{\hat{X}_{\tau_m , N}}\sum_{j=0}^{\tau -1}h(\hat{X}_{j,N})\\ &\le \frac{c}{N^2}\sum_{x \in \Delta_N}\frac{\Phi_N (x)}{|\Delta_N |}\sum_{m\ge 0}\big[\sup_{y\in \Delta_N}E_{\omega^N}^{y}(1-\frac{c}{N^2})^{\tau}\big]^{m}\cdot \sup_{y\in \Delta_N}E_{\omega^N}^{y} \sum_{j=0}^{\tau -1}h(\hat{X}_{j,N}). \end{align*} Since the function $f(x)=E_{\omega^N}^{x} \sum_{j=0}^{\tau -1}h(\hat{X}_{j,N})$ satisfies \begin{equation} \left\{ \begin{array}{rl} L_{\omega^N}f(x)=h(x), & \text{if } x\in\Delta_N\\ f(x)=0, & \text{if } x\in\partial \Delta_N, \end{array} \right. \end{equation} we can apply the maximum principle (Theorem \ref{CMP}) and get $$\sup_{y\in \Delta_N}E_{\omega^N}^{y}\sum_{j=0}^{\tau -1}h(\hat{X}_{j,N}) \le C N^2 \lVert \frac{h}{\varepsilon} \rVert _{\Delta_N ,d}.$$ This, together with Lemma \ref{CLTE} and $\sum_{x \in \Delta_N}\Phi_N (x)/|\Delta_N |=1$, yields $$\lVert \Phi_N \cdot h\rVert _{\Delta_N , 1} \le C \lVert \frac{h}{\varepsilon}\rVert _{\Delta_N ,d}.$$ Hence by the duality of norms, \begin{equation*} \lVert \Phi_N \varepsilon\rVert _{\Delta_N , \beta}=\sup_{\norm{h/\varepsilon}_{\Delta_N, d}=1}\norm{ \Phi_N h} _{\Delta_N, 1}\le C . \quad \mbox{\qed} \end{equation*} \noindent{\it Proof of (\ref{CPhi0}) under the assumption of Theorem \ref{CLT1}(i) :}\\ Assume that \begin{equation}\label{asm} \mathrm{E}\varepsilon(o)^{-p}< \infty \mbox{ for some } p>d. \end{equation} Take $\alpha=(1-1/d+1/p)^{-1}$. We use H\"older's inequality and Theorem \ref{CPhi} to get \[ \lVert \Phi_N\rVert _{\Delta_N,\alpha}\le \lVert \Phi_N \varepsilon\rVert _{\Delta_N , \beta}\lVert \varepsilon^{-1}\rVert _{\Delta_N, p} \le C \lVert \varepsilon^{-1}\rVert _{\Delta_N, p} . \] By the multidimensional ergodic theorem, \begin{equation*} \mathop{\rm lim}_{N \to \infty}\lVert \varepsilon^{-1}\rVert _{\Delta_N, p}=(E \varepsilon(o)^{-p})^{\frac{1}{p}}<\infty, \quad P\mbox{- }a.s. \qed \end{equation*} \begin{remark} Without the assumption (\ref{asm}), the conclusion \eqref{CPhi0} may fail. To see the difficulty, let $$A=A(\omega, \varepsilon_0)=\{x:\min_i \omega(x, e_i)<\varepsilon_0\}.$$ By (\ref{CPhi1}) we have $$\lVert \Phi_N 1_{A^c}\rVert _{\Delta_N, \beta}\le \lVert \Phi_N \frac{\varepsilon}{\varepsilon_0}\rVert _{\Delta_N, \beta} \le \frac{C}{\varepsilon_0}.$$ In order to proceed as before, we need to show that $\varlimsup_{N\to\infty}\lVert \Phi_N 1_{A}\rVert _{\Delta_N, \alpha} \le C$ for some $1 <\alpha \le \beta$ . As Bouchaud's trap model \cite{Bou,BAC} shows, this is not always the case. However, if $P\{\max_{|e|=1}\omega(o,e)\ge\xi_0\}=1$, then for $x\in A$, we have, using that the environment is balanced, some control of $\Phi_N (x)$ by $\Phi_N|_{A^c}$ (see Lemma \ref{CPhicontrol}). Further, in the iid case, $A$ corresponds to a `site percolation' model, whose cluster sizes can be estimated. We will show in the next section that these properties lead to a proof of (\ref{CPhi0}) in the iid setup, without moment assumptions. \end{remark} \section{A percolation estimate and proof of Theorem \ref{CLT2}(i)}\label{SePercE} In this section we consider the RWRE in the iid setting where $\max_{|e|=1}\omega(x,e)\ge \xi_0$ for all $x\in\mathbb{Z}^d$ and all $\omega\in\Omega$. We begin by introducing some terminology.\\ The \textit{$l^1$-distance} (graph distance) from $x$ to $y$ is defined as $$d(x,y)=|x-y|_1=\sum_{i=1}^{d}|x_i-y_i|.$$ Note that $|x|_{\infty}\le |x|_1 \le d|x|_{\infty}$. In an environment $\omega$, we say that a site $x$ is \textit{open}(\textit{closed}) if $\min_i \omega(x, e_i)<\varepsilon_0 (\ge\varepsilon_0, resp.)$ and that an edge of $\mathbb{Z}^d$ is open if its endpoints are open. Here $\varepsilon_0>0$ is a constant whose value is to be determined. An edge is called closed if it is not open. Let $A=A(\omega)$ denote the subgraph of $\mathbb{Z}^d$ obtained by deleting all closed edges and closed sites. We call $A(\omega)$ a \textit{site percolation} with parameter $p=p(\varepsilon_0)=P\{\min_i \omega(x, e_i)< \varepsilon_0\}$. A \textit{percolation cluster} is a connected component of $A$. (Although here a percolation cluster is defined as a graph, we also use it as a synonym for its set of vertices.) The $l^1$ diameter of a percolation cluster $B$ is defined as $l(B)=\sup_{x\in B, y\in \partial B}d(x,y)$. For $x\in A$, let $A_x$ denote the percolation cluster that contains $x$ and let $l_x$ denote its diameter. Set $A_x=\emptyset$ and $l_x=0$ if $x\notin A$. We let $\varepsilon_0$ be small enough such that $l_x<\infty$ for all $x\in \mathbb{Z}^d$.\\ We call a sequence of sites $(x^1, \cdots, x^n)$ a \textit{path} from $x$ to $y$ if $x^1=x$, $x^n=y$ and $|x^j-x^{j+1}|=1$ for $j=1,\cdots, n-1$. Let $$\square=\{(\kappa_1,\cdots, \kappa_d)\in \mathbb{Z}^d: \kappa_i=\pm 1\}.$$ We say that a path $\{x^1, \cdots, x^n\}$ is a \textit{$\kappa$-path}, $\kappa\in \square$, if $$\omega(x^j, x^{j+1}-x^j)\ge\xi_0$$ and $\kappa_i(x^{j+1}-x^j)_i\ge 0$ for all $i=1,\cdots, d$ and $j=1,\cdots, n-1$. Observing that for each site there exist at least two neighbors (in opposite directions) to whom the transition probabilities are $\ge \xi_0$, we have the following property concerning the structure of the balanced environment: \begin{itemize} \item For any $x\in A$ and any $\kappa\in \square$, there exists a $\kappa$-path from $x$ to some $y\in\partial A_x$, and this path is contained in $\bar{A}_x$. \end{itemize} This property gives us a useful inequality. \begin{lemma}\label{CPhicontrol} For $x\in A\cap \Delta_N$, if $l_x \le N$, then \begin{equation}\label{CLTf7} \Phi_N (x) \le \xi_0 ^{-l_x} \sum_{y \in \partial A_x \cap \Delta_N} \Phi_N (y). \end{equation} \end{lemma} {\it Proof:}~ Suppose that $A_x\neq\emptyset$ (otherwise the proof is trivial). Since $l_x \le N$, $\bar{A}_x\subset \Delta_N (x)$. Note that at least one of the $2^d$ corners of $\Delta_N (x)$ is contained in $\Delta_N$. Without loss of generality, suppose that $v=x+(N,\cdots,N)\subset \Delta_N$. Then there is a $(1,\cdots, 1)$-path in $\bar{A}_x$ from $x$ to some $y\in \partial A_x\cap \Delta_N$, as illustrated in the following figure: \begin{center} \begin{tikzpicture}[scale=.2]{centered} \fill[gray!30!white][rotate=45] (10, 1.5) ellipse (4.5 and 4.3); \draw[step=1, gray, very thin] (1.5,3.5) grid (10.5, 12.5); \draw[dash pattern=on 2pt off 3pt on 4pt off 4pt] (-5,-5) rectangle (15, 15) node[right=1pt]{$v$}; \draw (2,4) rectangle (22, 24); \draw[thick] (5,5)--(5,6)--(7,6)--(7,7)--(8,7)--(8,10)--(9,10)--(9,11)--(10,11)node[right=1pt]{$y$}; \draw[left=1pt] (5,5) node{$x$}; \draw (3,10) node[right=1pt]{$A_x$}; \draw (23, 14) node[right=1pt]{$\Delta_N$}; \draw (-5, 5) node[left=1pt]{$\Delta_N(x)$}; \end{tikzpicture} \end{center} Recalling that $\Phi_N$ is the invariant measure for $\{\hat{X}_{n,N}\}$ defined in Section 1, we have \begin{align*} \Phi_N (y) &= \sum_{z\in \Delta_N} \Phi_N (z) P_{\omega^N}^{d(x,y)} (\hat{z}, \hat{y})\\ &\ge \Phi_N (x) P_{\omega^N}^{d(x,y)}(\hat{x}, \hat{y}) \ge \Phi_N (x) \xi_0^{l_x}. \end{align*} Here $P^m_{\omega^N} (\hat{z}, \hat{y})$ denotes the $m$-step transition probability of $\{\hat{X}_{n,N}\}$ from $\hat{z}$ to $\hat{y}$. \qed Let $S_n=\{x: |x|_{\infty}=n\}$ denote the boundary of $\Delta_n$. Let $x\to y$ be the event that $y\in \bar{A}_x$ and $o\to S_n$ be the event that $o\to x$ for some $x\in S_n$. The following theorem, which is the site percolation version of the combination of Theorems 6.10 and 6.14 in \cite{GG}, gives an exponential bound on the diameter of the cluster containing the origin, when $p$ is small. \begin{theorem}\label{perc} There exists a function $\phi(p)$ of $p=p(\varepsilon_0)$ such that $$P\{o\to S_n\}\le C n^{d-1} e^{-n\phi(p)}$$ and $\mathop{\rm lim}_{p\to 0}\phi(p)=\infty$. \end{theorem} Let $A_x(n)$ denote the connected component of $A_x\cap \Delta_n(x)$ that contains $x$ and set \[q_n=P\{o\to S_n\}.\] The proof of Theorem \ref{perc} will proceed by showing some (approximate) subadditivity properties of $q_n$. We thus recall the following subadditivity lemma: \begin{lemma} \label{Clem-subadd} If a sequence of finite numbers $\{b_k: k\ge 1\}$ is subadditive, that is, $b_{m+n}\le b_m +b_n \mbox{ for all m,n}$, then $\mathop{\rm lim}_{k\to\infty}b_k/k=\inf_{k\in\mathbb{N}} b_k/k$. \end{lemma} \noindent{\it Proof of Theorem \ref{perc}:} We follow the proof given by Grimmett in \cite{GG} in the bond percolation case. By the BK inequality (\cite{GG}, pg. 38), $$q_{m+n}\le \sum_{x\in S_m} P\{o\to x\} P\{x\to x+S_n\}.$$ But $P\{o\to x\}\le q_m$ for $x\in S_m$ and $P\{x\to x+S_n\}=q_n$ by translation invariance. Hence we get \begin{equation}\label{Csub1} q_{m+n}\le |S_m|q_m q_n. \end{equation} By exchanging $m$ and $n$ in (\ref{Csub1}), \begin{equation}\label{Csub11} q_{m+n}\le |S_{m\wedge n}|q_m q_n. \end{equation} On the other hand, let $U_x$ be the event that $x\in \overline{A_o(m)}$ and let $V_x$ be the event that $\overline{A_x(n)}\cap S_{m+n}\neq\emptyset$. We use the FKG inequality (\cite{GG}, pg. 34) to find that $$q_{m+n}\ge P\{U_x\}P\{V_x\}\quad \mbox{ for any $x\in S_m$}.$$ However, $\sum_{x\in S_m}P\{U_x\}\ge q_m$, which implies that $$\max_{x\in S_m}P\{U_x\}\ge \frac{q_m}{|S_m|} .$$ Let $\gamma_n=P\{\overline{A_o(n)}\cap\{x:x_1=n\}\neq\emptyset\}$, then $P\{V_x\}\ge\gamma_n$. Moreover, $\gamma_n\le q_n\le 2d\gamma_n$. Hence \begin{equation*} q_{m+n}\ge \frac{q_m q_n}{2d |S_m|}, \end{equation*} and then \begin{equation}\label{Csub2} q_{m+n}\ge \frac{q_m q_n}{2d |S_{m\wedge n}|}. \end{equation} Note that $|S_m|\le C_d m^{d-1}$. Letting $$b_k =\log q_k +\log C_d +(d-1)\log (2k),$$ one checks using (\ref{Csub11}) that the sequence $\{b_k\}$ is subadditive. Similarly by (\ref{Csub2}), $\{-\log q_k +\log (2d C_d) +(d-1)\log (2k)\}$ is subadditive. Thus, using Lemma \ref{Clem-subadd}, $$\phi (p):=-\mathop{\rm lim}_{k\to\infty}\frac{1}{k}\log q_k$$ exists and \begin{equation}\label{Csub3} \log q_k +\log C_d +(d-1)\log (2k)\ge -k\phi (p), \end{equation} \begin{equation}\label{Csub4} -\log q_k +\log (2d C_d) +(d-1)\log (2k)\ge k\phi(p). \end{equation} The first part of the theorem follows simply from (\ref{Csub4}), and the second by noting that with $p\downarrow 0$ in (\ref{Csub3}) we have $q_k \downarrow 0$ and then $\phi (p)\to \infty$. \qed \begin{remark} It follows from Theorem \ref{perc} that \begin{equation}\label{Clo} P\{l_o \ge n\} \le P\{o\to S_{\lfloor n/2d\rfloor}\}\le C e^{\phi(p)} n^{d-1} e^{-n\phi(p)/2d}. \end{equation} With (\ref{Clo}) and the Borel-Cantelli lemma, one concludes that, P-almost surely, $l_x\le N$ is true for all $x\in \Delta_N$ when $N$ is sufficiently large and $p$ is such that $\phi(p)>0$. Hence the inequality (\ref{CLTf7}) holds for all $x\in\Delta_N$ when $N$ is large. \end{remark} \noindent{\it Proof of (\ref{CPhi0}) under the assumption of Theorem \ref{CLT2}(i):} By H\"older's inequality, $$\frac{1}{|\Delta_N|}\sum_{y\in \partial A_x \cap \Delta_N} \Phi(y)\le \lVert \Phi_N 1_{\partial A_x}\rVert _{\Delta_N, \beta} \big(\frac{|\partial A_x|}{|\Delta_N|}\big)^{1-1/\beta} ,$$ so when $N$ is large enough we have by Lemma \ref{CPhicontrol} that for any $x\in A\cap \Delta_N$, \begin{equation}\label{CLTf8} \Phi_N (x)\le \xi_0^{-l_x} |\partial A_x|^{1-1/\beta } |\Delta_N |^{1/\beta} \lVert \Phi_N 1_{\partial A_x}\rVert _{\Delta_N, \beta}. \end{equation} Hence for any $\alpha \in (1, \beta)$, \begin{align*} &\lVert \Phi_N 1_A\rVert _{\Delta_N,\alpha}^\alpha\\ &\le \frac{1}{|\Delta_N|}\sum_{x\in A\cap\Delta_N}\big(\xi_0^{-l_x} |\partial A_x|^{1-1/\beta} |\Delta_N |^{1/\beta}\lVert \Phi_N 1_{\partial A_x}\rVert _{\Delta_N, \beta} \big)^\alpha\\ &\le \left[\frac{1}{|\Delta_N|}\sum_{x\in A\cap\Delta_N} \big(\xi_0^{-l_x} |\partial A_x|^{1-1/\beta}|A_x|^{1/\beta})^{\alpha (\beta/\alpha)'}\right] ^{1-\alpha/\beta}\\ &\qquad\times \left[\frac{1}{|\Delta_N|}\sum_{x\in A\cap\Delta_N} \big(\frac{|\Delta_N|^{1/\beta}\lVert \Phi_N 1_{\partial A_x}\rVert _{\Delta_N, \beta}} {|A_x|^{1/\beta}}\big)^{\beta} \right]^{\alpha/\beta}\\ &= \left[\frac{1}{|\Delta_N|}\sum_{x\in A\cap\Delta_N} \big(\xi_0^{-l_x} |\partial A_x|^{1-1/\beta}|A_x|^{1/\beta}\big)^{\alpha\beta/(\beta -\alpha)}\right] ^{1-\alpha/\beta}\\ &\qquad\times \left(\sum_{x\in A\cap\Delta_N} \frac{\lVert \Phi_N 1_{\partial A_x}\rVert _{\Delta_N, \beta}^{\beta}}{|A_x|} \right)^{\alpha/\beta}, \end{align*} where we used (\ref{CLTf8}) in the first inequality and H\"older's inequality in the second. Observe that \begin{equation}\label{CPhiaverage} \sum_{x\in A\cap\Delta_N} \dfrac{\lVert \Phi_N 1_{\partial A_x}\rVert _{\Delta_N, \beta}^{\beta}}{|A_x|} \le \sum_{i=1}^n \lVert \Phi_N 1_{\partial A_i}\rVert _{\Delta_N, \beta}^{\beta} \le 2d \lVert \Phi_N 1_{\partial A}\rVert _{\Delta_N, \beta}^{\beta} \le C\varepsilon_0^{-\beta}, \end{equation} where $A_1, \cdots, A_n$ are different clusters that intersect with $\Delta_N$. On the other hand, the multidimensional ergodic theorem gives \begin{align}\label{percergodic} &\mathop{\rm lim}_{N\to\infty}\frac{1}{|\Delta_N|}\sum_{x\in A\cap\Delta_N} \big(\xi_0^{-l_x} |\partial A_x|^{1-1/\beta}|A_x|^{1/\beta}\big)^{\alpha\beta/(\beta -\alpha)}\nonumber\\ &= E \big(\xi_0^{-l_o} |\partial A_o|^{1-1/\beta}|A_o|^{1/\beta}\big)^{\alpha\beta/(\beta -\alpha)} \le C E \big(\xi_0^{-l_o} l_o^d\big)^{\alpha\beta/(\beta -\alpha)} \quad \mbox{P-a.s.,} \end{align} which by (\ref{Clo}) is finite when $\varepsilon_0$ is small. \qed \section{Transience in general ergodic environments}\label{SeTran} In this section we will prove (ii) of Theorem \ref{CLT1} by an argument similar to that in \cite{ZO}. The main differences in our method are that we use a stronger control of the hitting time (Lemma \ref{Ctau}), and that we apply a mean value inequality (Theorem \ref{Cmvi}) instead of the discrete Harnack inequality used in \cite{ZO}. \begin{lemma}\label{Ctau} Let $\{X_n\}$ be a random walk in a balanced environment $\omega$ such that $\omega(x,o)=0$ for all $x$. For any $r>0$, define $\tau=\tau(r)= \inf\{n: |X_n|>r\}$. Then $E_\omega^o \tau\le (r+1)^2$. \end{lemma} {\it Proof:}~ Observe that $\{|X_n|^2-n\}$ is a (quenched) martingale with respect to $\{\mathcal{F}_n=\sigma(X_1,\cdots,X_n)\}$. Thus by optional stopping, $0= E_\omega^o[|X_\tau|^2-\tau]\le (r+1)^2-E_\omega^o\tau$. \qed To prove Theorem \ref{CLT1}(ii), we shall make use of the following mean-value inequality, which is a modification of Theorem 3.1 in \cite{KT}. Let $B_{r}(z)=\{x\in\mathbb{Z}^d: |x-z| <r\}$. We shall also write $B_r (o)$ as $B_r$; recall the definition of $L_\omega$ in \eqref{operator}. \begin{theorem}\label{Cmvi} For any function $u$ on $\bar B_R (x_0)$ such that $$L_{\omega} u =0 , \quad x \in B_R (x_0)$$ and any $\sigma\in (0,1)$, $0<p\le d$, we have \[ \max_{B_{\sigma R}(x_0)}u\le C \lVert \frac{u^+}{\varepsilon^{d/p}}\rVert _{B_R (x_0), p}, \] where $C$ depends on $\sigma$, $p$ and $d$. \end{theorem} We postpone the proof of Theorem \ref{Cmvi} to the next section, and now demonstrate Theorem \ref{CLT1}. \noindent{\it Proof of Theorem \ref{CLT1}(ii):} Note that the transience of the random walk would not change if we considered the walk restricted to its jump times. That is, the transience or recurrence of the random walk in an environment $\omega$ is the same as in an environment $\tilde \omega$, where $\tilde \omega$ is defined by $\tilde{\omega}(x,e)=\omega (x,e)/(1-\omega(x,o))$. Therefore, in the sequel we assume $\omega(x,o)=0$ for all $x$ and almost all $\omega$. Let $K$ be any constant that is at least 3. We denote $B_{K^i}(x)$ by $B^i(x)$ and define $\tau_i:=\inf \{n: |X_n|> K^i\}$. Our approach is to bound the (annealed) expected number of visits to the origin by the walk; this requires some a-priori bounds on the moments of $\varepsilon(o)^{-1}$.\\ For any $z\in\partial B^i$ , $y\in B^{i-1}$, noting that $E_\omega^x (\mbox{\# visits at $y$ before $\tau_{i+2}$}):=v(x)$ satisfies $L_\omega v (x)=0$ for $x\in B^{i+2}\setminus\{y\}$, we have that, for $p\in(0,d]$, \begin{align}\label{CLTf11} & E_{\theta^{y}\omega}^{z} (\mbox{ \# visits at $o$ before $\tau_{i+1}$}) \nonumber\\ &\le E_{\omega}^{z+y}(\mbox{\# visits at $y$ before $\tau_{i+2}$})\nonumber\\ &\le \max_{x\in B^{i-1}(z)}E_{\omega}^{x}(\mbox{\# visits at $y$ before $\tau_{i+2}$})\nonumber\\ &\le C\Bnorm{\frac{E_{\omega}^{x}( \mbox{\# visits at $y$ before $\tau_{i+2}$})}{\varepsilon_\omega (x)^{d/p}} } _{B_{2K^{i-1}}(z), p}\nonumber\\ &\le C\Bnorm{ \frac{E_{\omega}^{x} (\mbox{\# visits at $y$ before $\tau_{i+2}$})}{\varepsilon_\omega (x)^{d/p}} } _{B^{i+2}, p}, \end{align} where we used Theorem \ref{Cmvi} in the third inequality. Take $p=d/q$ (without loss of generality, we always assume that $q< d$). Then, by (\ref{CLTf11}) and Lemma \ref{Ctau}, \begin{align}\label{CLTf12} & \sum_{y\in B^{i-1}}E_{\theta^{y}\omega}^o (\mbox{ \# visits at $o$ in $[\tau_i,\tau_{i+1})$})\nonumber\\ &\le C\sum_{y\in B^{i-1}} \left[ \frac{1}{|B^{i+2}|}\sum_{x\in B^{i+2}} \frac{ E_{\omega}^{x} (\mbox{\# visits at $y$ before $\tau_{i+2}$})^{d/q}} {\varepsilon_\omega (x)^d} \right]^{q/d}\nonumber\\ &\le C K^{-iq}\sum_{y\in B^{i-1}}\sum_{x\in B^{i+2}} \frac{E_{\omega}^{x}(\mbox{\# visits at $y$ before $\tau_{i+2}$})} {\varepsilon_\omega (x)^q} \nonumber\\ &= C K^{-iq}\sum_{x\in B^{i+2}} \frac{E_{\omega}^{x}(\mbox{\# visits at $B^{i-1}$ before $\tau_{i+2}$})} {\varepsilon_\omega (x)^q}\nonumber\\ &\le C K^{-iq}\sum_{x\in B^{i+2}}\frac{E_{\omega}^{x}\tau_{i+2}}{\varepsilon_\omega (x)^q}\nonumber\\ &\le C K^{(2-q)i} \sum_{x\in B^{i+2}} \varepsilon_\omega(x)^{-q}. \end{align} Taking expectations and using translation invariance, we have \begin{equation*} \mathbb{E}^o(\mbox{\# visits at $o$ in $[\tau_i, \tau_{i+1})$}) \le C K^{(2-q)i} E \varepsilon^{-q}. \end{equation*} Therefore, if $E \varepsilon^{-q}<\infty$ for some $q>2$ , then \begin{equation*} \mathbb{E}^o (\mbox{\# visits at $o$})\le C E \varepsilon^{-q} \sum_{i=1}^{\infty}K^{(2-q)i}<\infty . \end{equation*} This proves Theorem \ref{CLT1}(ii) for $\{\Omega, P\}$ such that $\omega (x,o)=0$ for all $x$ and almost all $\omega$. As mentioned earlier, the general case follows by replacing $\varepsilon$ with $\varepsilon/(1-\omega(o,o))$.\qed\\ \begin{remark} It is natural to expect that arguments similar to the proof of the invariance principle also work for proving the transience in the iid case. Namely, one may hope to control $P_\omega^x\{\mbox{visit $o$ in $[\tau_i,\tau_{i+1})$}\}$ using some mean value inequality (like Theorem \ref{Cmvi}), and to use percolation arguments to handle ``bad sites'' where the ellipticity constant $\varepsilon$ is small. This suggests considering walks that jump from bad sites to good sites. In \cite{KT2}, Kuo and Trudinger proved a maximum principle and mean value inequality for balanced operators in general meshes, which may be applied to balanced walks with possibly big jumps. However, their estimates, in the presence of a small ellipticity constant, are not strong enough. To overcome this issue, we will prove a modified maximum principle that involves only big exit probabilities, and then use it to prove the transience in the i.i.d case with no moment assumptions. \end{remark} \section{Transience in iid environments}\label{SeTriid} In this section we prove a modified maximum principle for balanced environments. We then prove Theorem 2(ii) using the corresponding mean value inequality (Theorem \ref{Cmvi2}) and percolation arguments. \subsection{Difference operators}\label{SeTriid1} Following \cite{KT2}, we introduce general difference operators. Let $a$ be a nonnegative function on $\mathbb{Z}^d\times \mathbb{Z}^d$ such that for any $x$, $a(x,y)> 0$ for only finitely many $y$. Define the linear operator $L_a$ acting on the set of functions on $\mathbb{Z}^d$ by \[L_a f(x)=\sum_y a(x,y)(f(y)-f(x)).\] We say that $L_a$ is \textit{balanced} if \begin{equation}\label{CLTe7} \sum_y a(x,y)(y-x)=0. \end{equation} Throughout this section we assume that $L_a$ is a probability operator, that is, \[\sum_y a(x,y)=1.\] For any finite subset $E\subset\mathbb{Z}^d$, define its boundary \[E^b=E^b(a)=\{y\notin E: a(x,y)>0 \text{ for some } x\in E \}, \] and set \begin{equation} \label{CLTeq-ofernew} \tilde{E}=E\cup E^b. \end{equation} Define the upper contact set of $u$ at $x\in E$ as $$I_u(x)=I_u(x,E,a)=\{s\in\mathbb{R}^d: u(x)-s\cdot x\ge u(z)-s\cdot z \text{ for all }z\in\tilde{E}\}.$$ Set \begin{align*} &h_x=h_x(a)=\max_{y: a(x,y)>0}\abs{x-y},\\ &b(x)=\sum_y a(x,y)(y-x), \mbox{ and }b_0=\sup|b|. \end{align*} Note that $b_0=0$ when $L_a$ is balanced. The following lemma is useful in the proofs of various mean value inequalities. It is similar to Theorem 2.2 in \cite{KT2}, except that the proof in \cite{KT2} contains several unclear passages, e.g., in the inequality above (2.23) in \cite{KT2}, and so we provide a complete proof. Throughout, we set $u^+=u\vee 0$. \begin{lemma}\label{Cmvilemma} Fix $R>0$. Let $\eta(x)=\eta_R(x):=(1-\abs{x}^2/R^2)^\beta 1_{|x|<R}$ be a function on $\mathbb{R}^d$. For any function $u$ on $B_R$ such that $L_a u\ge 0$ in $B_R$ and any $\beta\ge 2$, we let $v=\eta u^+$. Then, for any $x\in B_R$ with $I_v (x)=I_v(x, B_R, a)\neq\emptyset$, \[L_a v(x)\ge -C(\beta, b_0R) \eta^{1-2/\beta}R^{-2} h_x^2 u^+,\] where $C(\beta,b_0R)$ is a constant that depends only on $\beta$ and $b_0R$. \end{lemma} {\it Proof:}~ We only need to consider the nontrivial case where $v\not\equiv 0$. For $s=s(x)\in I_v(x)\neq \emptyset$, recalling the definition of $I_v$ one has $$|s|\le 2v(x)/(R-|x|).$$ Note that $I_v(x)\neq\emptyset$ implies $u(x)> 0$. If further $R^2-|x|^2\ge 4R \abs{x-y}$ , computations as in \cite[pg. 426]{KT2} reveal that \begin{align} 2^{-\beta} &\le \frac{\eta(y)}{\eta(x)}\le 2^\beta,\label{Cfirst}\\ \abs{\eta(x)-\eta(y)} &\le \beta 2^\beta R^{-1} \eta(x)^{1-1/\beta}|x-y|,\label{Csecond}\\ \abs{\eta(x)-\eta(y)-\nabla\eta(x)(x-y)} &\le \beta (\beta-1)2^{\beta} R^{-2}\eta(x)^{1-2/\beta}|x-y|^2,\label{Cthird}\\ |s| &\le 4 \eta^{1-1/\beta}R^{-1}u,\label{Cfourth} \end{align} where \begin{equation}\label{Afifth} \nabla\eta=-2\beta x R^{-2}\eta^{1-1/\beta} \end{equation} is the gradient of $\eta$. Following \cite{KT2}, we set $w(z)=v(z)-s\cdot (z-x)$. By the definition of $s$, we have $w(x)\ge w(z)$ for all $z\in \tilde E$. Then \begin{align}\label{A*6} v(x)-v(y)=\frac{\eta(x)}{\eta(y)}(v(x)-v(y)) &+\frac{\eta(y)-\eta(x)}{\eta(y)}s(x-y)\\ &+\frac{\eta(y)-\eta(x)}{\eta(y)}(w(x)-w(y)).\nonumber \end{align} Consider first $x$ such that $R^2-|x|^2\ge 4Rh_x$. By (\ref{Csecond}), for any $y$ such that $a(x,y)>0$, \begin{equation}\label{A*7} \frac{\eta(y)-\eta(x)}{\eta(y)}(w(x)-w(y)) \le \beta 2^\beta R^{-1}h_x\eta(x)^{-1/\beta}\frac{\eta(x)}{\eta(y)}(w(x)-w(y)). \end{equation} Since \begin{align*} &\sum_y a(x,y)\frac{\eta(x)}{\eta(y)}(w(x)-w(y))\\ &=\sum_y a(x,y) \Big[\frac{\eta(x)}{\eta(y)}\big(v(x)-v(y)\big) +\frac{\eta(x)-\eta(y)}{\eta(y)}s (y-x)\Big]+s\cdot b(x), \end{align*} by \eqref{A*6}, \eqref{A*7} and noting $R^{-1}\eta^{-1/\beta}h_x\le 1/4$, we obtain \begin{align}\label{A*8} &\sum_y a(x,y)(v(x)-v(y))\nonumber\\ &\le (1+\beta 2^\beta R^{-1}h_x\eta(x)^{-1/\beta}) \sum_y a(x,y)\Big[\frac{\eta(x)}{\eta(y)}\big(v(x)-v(y)\big)+\frac{\eta(x)-\eta(y)}{\eta(y)}s (y-x)+s\cdot b(x)\Big] \nonumber\\ &\le \beta 2^{\beta-1}\big[\sum_y a(x,y) \frac{\eta(x)}{\eta(y)}\big(v(x)-v(y)\big)+ 4(\beta 2^{2\beta}+b_0R)\eta^{1-2/\beta}R^{-2}h_x^2u\big], \end{align} where we used \eqref{Cfirst}, \eqref{Csecond} \eqref{Cfourth} in the last inequality. Moreover, recalling that $u(x)>0$ (because $I_v(x)\neq\emptyset$), \begin{align}\label{A*9} & \sum_y a(x,y) \frac{\eta(x)}{\eta(y)}\big(v(x)-v(y)\big)\nonumber\\ &= \sum_y a(x,y)\left [\eta(x)\big(u(x)-u^+(y)\big)+\big(\eta(x)-\eta(y)\big)u(x)+\frac{(\eta(x)-\eta(y))^2}{\eta(y)}u(x)\right]\nonumber\\ &\stackrel{a\geq 0}{\le} -\eta(x)L_a u(x)+\sum_y a(x,y) \left[\big(\eta(x)-\eta(y)\big)u(x)+\frac{(\eta(x)-\eta(y))^2}{\eta(y)}u(x)\right]\nonumber\\ &\stackrel{L_au\ge 0}{\le} \sum_y a(x,y) \left[\big(\eta(x)-\eta(y)-\nabla\eta(x)(x-y)\big)u(x)+\frac{(\eta(x)-\eta(y))^2}{\eta(y)}u(x)\right]\nonumber\\ &\qquad\qquad-\nabla\eta(x)b(x)u(x)\nonumber\\ &\le (2^{3\beta+1}+b_0R)\beta^2\eta^{1-2/\beta}h_x^2R^{-2}u, \end{align} where we used (\ref{Cfirst}), (\ref{Csecond}), (\ref{Cthird}) and (\ref{Afifth}) in the last inequality. Hence, by (\ref{A*8}) and (\ref{A*9}), we conclude that \begin{equation*} -L_a v \le (2^{3\beta+1}+b_0R)\beta^32^\beta \eta^{1-2/\beta}R^{-2}h_x^2 u \end{equation*} holds in $\{x: R^2-|x|^2\ge 4Rh_x, I_v(x)\neq\emptyset\}$. On the other hand, if $R^2-|x|^2<4Rh_x$, then $\eta^{1/\beta}\le 4h_x/R$. Thus by the fact that $u(x)>0$, we have $-L_a v\le v(x)\le 16\eta^{1-2/\beta} R^{-2} h_x^2 u$. \qed\\ \noindent{\it Proof of Theorem \ref{Cmvi}:} Since $L_\omega$ is a balanced operator ($b_0=0$) and $h_x=1$ in this case, by the above lemma, \[L_\omega v\ge -C(\beta) \eta^{1-2/\beta}R^{-2} u\] for $x\in B_R$ such that $I_u(x)\neq\emptyset$, where $C(\beta)$ depends only on $\beta$. Applying Theorem \ref{CMP} to $v$ and taking $\beta=2d/p\ge 2$, we obtain \begin{align*} \max_{B_R} v &\le C \Bnorm{ \eta^{1-2/\beta} \frac{u^+}{\varepsilon}} _{B_R, d} =C\Bnorm{ v^{1-p/d}\frac{(u^+)^{p/d}}{\varepsilon}}_{B_R, d}\\ &\le C(\max_{B_R} v)^{1-p/d}\Bnorm{\frac{u^+}{\varepsilon^{d/p}}} _{B_R, p}^{p/d}. \end{align*} Hence \[ \max_{B_R} v \le C \Bnorm{\frac{u^+}{\varepsilon^{d/p}}} _{B_R, p}, \] and then \[ \max_{B_{\sigma R}}u\le (1-\sigma^2)^{-2d/p}\max_{B_{\sigma R}}v \le C(\sigma, p, d) \Bnorm{\frac{u^+}{\varepsilon^{d/p}}} _{B_R, p}. \text{\qed} \] \subsection{A new maximum principle and proof of Theorem \ref{CLT2}(ii)} For any fixed environment $\omega\in\Omega$, let $\varepsilon_0>0$ be a constant to be determined, and define site percolation as in Section \ref{SePercE}. Recall that for $x\in\mathbb{Z}^d$, $A_x$ is the percolation cluster that contains $x$ and $l_x$ is its $l^1$-diameter. As mentioned in the introduction, the transience would not change if we considered the walk restricted to its jump times. Without loss of generality, we assume that $\omega(x,o)=0$ for all $x$, $P$-almost surely. Recall the definition of $\square$ and $\kappa$-path for $\kappa\in\square$ in Section \ref{SePercE}. Note that under our assumption, $\max_i \omega(x, e_i)\ge 1/2d$, so we take $\xi_0=1/2d$ in the definition of $\kappa$-paths. For each $\kappa\in\square$, we pick a site $y_\kappa=y(x, \kappa)\in \partial A_x$ such that \[d(x, y_\kappa)=\max_{\substack{y: \exists \text{ $\kappa$-path in $\bar{A}_x$ }\\\text{ from $x$ to $y$} } } d(x,y)\] and let $\Lambda_x\subset \bar{A}_x$ be the union of (the points of the) $\kappa$-paths from $x$ to $y_\kappa$ over all $\kappa\in\square$. From the definition of $y_\kappa$ one can conclude that \begin{itemize} \item For any $q\in\mathbb{R}^d$, we pick a $\kappa=\kappa_q\in \square$ such that \[q_j \kappa_j\le 0 \text{ for all }j=1,\cdots, d.\] Then $(y_\kappa-x)_j q_j\le 0$ for all $j=1,\cdots, d$. Moreover, for $i\in\{1,\cdots, d\}$, $q_i>0$ implies $y_\kappa-e_i\notin \Lambda_x$, and $q_i<0$ implies $y_\kappa+e_i\notin \Lambda_x$. \end{itemize} In the sequel we let $\tau_{\Lambda_x}=\inf\{n>0: X_n\notin \Lambda_x\}$ and \[a(x,y)=P_\omega^x \{ X_{\tau_{\Lambda_x}}=y \}. \] By the fact that $X_n$ is a (quenched) martingale, it follows that $L_a$ is a balanced operator. For the statement of the next theorem, recall the definition of $\tilde{E}$ in \eqref{CLTeq-ofernew}. \begin{theorem}\label{Cmp2} Let $E\subset\mathbb{Z}^d$ be bounded. Let $u$ be a function on $\tilde{E}$. If $L_a u(x)\ge -g(x)$ for all $x\in E$ such that $I_u(x)=I_u(x,E,a)\neq \emptyset$ , then $$\max_E u\le \frac{d\mathop{\rm diam} \tilde{E}}{\varepsilon_0} \bigg(\sum_{\substack{ x\in E\\I_u (x)\ne \emptyset } } \abs{g(x)(2d)^{l_x}}^d \bigg)^{\frac{1}{d}} +\max_{E^b}u .$$ \end{theorem} {\it Proof:}~ Without loss of generality, assume $g\ge 0$ and $$\max_E u=u(x_0)>\max_{E^b}u$$ for some $x_0\in E$. Otherwise, there is nothing to prove. For $ s\in \mathbb{R}^{d} $ such that $ |s|_{\infty} \le [u(x_0)-\max_{E^b}u]/(d\mathop{\rm diam} \tilde{E}) $, we have \[ u(x_0)-u(x) \ge s \cdot (x_0-x) \] for all $ x \in E^b $, which implies that $\max_{z\in\tilde{E}}u(z)-s\cdot z$ is achieved in $E$. Hence $ s \in \bigcup_{x \in E} I_{u}(x) $ and \begin{equation}\label{CLTe1} \left[-\dfrac{u(x_0)-\max_{E^b }u}{d\mathop{\rm diam} \tilde{E}}, \dfrac{u(x_0)-\max_{E^b}u}{d\mathop{\rm diam} \tilde{E}}\right]^{d} \subset \bigcup_{x\in E} I_{u}(x) . \end{equation} Further, if $s\in I_u(x)$, we set $$w(z)=u(z)-s(z-x).$$ Then $w(z)\le w(x)$ for all $z\in \tilde{E}$ and \begin{equation}\label{CLTe2} I_u(x)=I_w(x)+s. \end{equation} Since for any $q\in I_w(x)$, there is $\kappa=\kappa_q\in\square$ such that $$q_j(x-y_\kappa)_j\ge 0 \text{ for } j=1,\cdots, d,$$ we have $$ w(x)-w(y_\kappa\pm e_i)\ge q(x-y_\kappa\mp e_i)\ge \mp q_i.$$ Moreover, for any $i\in\{1,\cdots, d\}$, if $q_i>0$, then $y_\kappa-e_i\notin \Lambda_x$ and we have $w(x)-w(y_\kappa-e_i)\ge |q_i|$. Similarly, if $q_i<0$, then $y_\kappa+e_i\notin \Lambda_x$ and $w(x)-w(y_\kappa+e_i)\ge |q_i|$. We conclude that $$|q_i|\le \frac{\sum_{y} a(x, y)(w(x)-w(y))}{\min\limits_{\pm}\{a(x, y_\kappa\pm e_i)\}}.$$ On the other hand, from the construction of $\Lambda_x$ we obtain (noting that $y_\kappa\in\partial A_x$) $$a(x, y_\kappa\pm e_i)\ge (\frac{1}{2d})^{l_x} \varepsilon_0.$$ Hence, since $L_a$ is balanced, $$|q_i|\le \frac{(2d)^{l_x}}{\varepsilon_0}\sum_y a(x,y)(w(x)-w(y)) = \frac{(2d)^{l_x}}{\varepsilon_0}(-L_a u) \le \frac{(2d)^{l_x}}{\varepsilon_0} g$$ for all $i$. Therefore \begin{equation}\label{CLTe3} I_w(x)\subset [-(2d)^{l_x}\varepsilon_0^{-1} g, (2d)^{l_x}\varepsilon_0^{-1} g]^d. \end{equation} Combining (\ref{CLTe1}), (\ref{CLTe2}) and (\ref{CLTe3}), we conclude that \begin{equation*} \left(\dfrac{u(x_0)-\max_{E^b}u}{d\mathop{\rm diam} \tilde{E}}\right)^d\le \sum_{\substack{ x\in E\\I_u (x)\ne \emptyset } } \abs{g(x)(2d)^{l_x}\varepsilon_0^{-1}}^d. \quad\quad\quad\quad\quad \quad\quad\quad\quad\quad \mbox{\qed} \end{equation*} As with Theorem \ref{Cmvi}, we have a corresponding mean value inequality. \begin{theorem}\label{Cmvi2} For any function $u$ on $B_R$ such that $$L_a u=0, \quad x\in B_R$$ and any $\sigma\in (0,1)$, $0<p\le d$, we have \[ \max_{B_{\sigma R}} u\le C\big(\frac{\mathop{\rm diam} \tilde{B}_R}{\varepsilon_0 R}\big)^{d/p} \norm{[l_x^2 (2d)^{l_x}]^{d/p}u^+}_{B_R, p}, \] where $C$ depends on $\sigma, p$ and $d$. \end{theorem} {\it Proof:}~ By the same argument as in the proof of Theorem \ref{Cmvi}, Lemma \ref{Cmvilemma} and Theorem \ref{Cmp2} implies Theorem \ref{Cmvi2}. \qed\\ Having established Theorem \ref{Cmvi2}, we can now prove the transience of the random walks in balanced iid environment with $d\geq 3$. \noindent{\it Proof of Theorem \ref{CLT2}(ii)}: Let $K$ be any constant $\ge 4$ and define $B^i, \tau_i$ as in Section 5. Let $\Omega_i=\{\omega\in\Omega: l_x\le K^{i-1} \mbox{ for all $x\in B^{i+2}$}\}$. For any $\omega\in\Omega_i$, $z\in\partial B^i$, $y\in B^{i-1}$, noting that $P_\omega^x \{\mbox{visit $y$ before $\tau_{i+2}$}\}:=u(x)$ satisfies \[L_a u(x)=0\] for $x\in B_{2K^{i-1}}(z)$, by similar argument as in (\ref{CLTf11}) we have \begin{align*} & P_{\theta^y\omega}^z \{\mbox{visit $o$ before $\tau_{i+1}$}\}1_{\omega\in\Omega_i}\\ &\le \max_{x\in B^{i-1}(z)}P_\omega^x \{\mbox{visit $y$ before $\tau_{i+2}$}\}1_{\omega\in\Omega_i}\\ &\le C \varepsilon_0^{-d}\norm{ [l_x^2 (2d)^{l_x}]^d P_\omega^x \{\mbox{visit $y$ before $\tau_{i+2}$}\} }_{B_{2K^{i-1}}(z), 1}\\ &\le C \varepsilon_0^{-d}\abs{B^{i+2}}^{-1} \sum_{x\in B^{i+2}}l_x^{2d} (2d)^{dl_x} P_\omega^x \{\mbox{visit $y$ before $\tau_{i+2}$}\} , \end{align*} where in the second inequality, we applied Theorem \ref{Cmvi2} with $p=1$ and used the fact that $\mathop{\rm diam} \tilde{B}_{2K^{i-1}}\le 3K^{i-1}$ when $\omega\in\Omega_i$. Hence \begin{align}\label{CLTe5} &\sum_{y\in B^{i-1}}P_{\theta^y\omega}^o \{\mbox{visit $o$ in $[\tau_i, \tau_{i+1})$}\} 1_{\omega\in\Omega_i}\nonumber\\ &\le C \varepsilon_0^{-d}\abs{B^{i+2}}^{-1} \sum_{x\in B^{i+2}}l_x^{2d} (2d)^{dl_x} E_\omega^x (\mbox{\# visits at $B^{i-1}$ before $\tau_{i+2}$})\nonumber\\ &\stackrel{\text{Lemma }\ref{Ctau}}{\le} C \varepsilon_0^{-d}K^{(2-d)i}\sum_{x\in B^{i+2}}l_x^{2d} (2d)^{dl_x}. \end{align} Since \begin{align}\label{CLTe6} &\sum_{y\in B^{i-1}}P_{\theta^y\omega}^o \{\mbox{visit $o$ in $[\tau_i, \tau_{i+1})$}\}\nonumber\\ &\le \sum_{y\in B^{i-1}}P_{\theta^y\omega}^o \{\mbox{visit $o$ in $[\tau_i, \tau_{i+1})$}\}1_{\omega\in\Omega_i} +\abs{B^{i-1}}1_{\omega\notin\Omega_i}, \end{align} taking $P$-expectations on both sides of (\ref{CLTe6}) and using (\ref{CLTe5}) we get \[ \mathbb{P}^o \{\mbox{visit $o$ in $[\tau_i,\tau_{i+1})$}\} \le C \varepsilon_0^{-d}K^{(2-d)i}El_o^{2d}(2d)^{dl_o}+P\{\omega\notin\Omega_i\}. \] By (\ref{Clo}), we can take $\varepsilon_0$ to be small enough such that $El_o^{2d}(2d)^{dl_o}<\infty$ and $\sum_{i=1}^\infty P\{\omega\notin\Omega_i\}<\infty$. Therefore when $d\ge 3$, \[\sum_{i=1}^\infty \mathbb{P}^o \{\mbox{visit $o$ in $[\tau_i,\tau_{i+1})$}\}<\infty. \quad\quad\quad\quad\quad\quad \mbox{\qed}\] \section{Concluding remarks} While Bouchaud's trap model (see \cite{Bou,BAC}) provides an example of an (iid) environment where local traps can destroy the invariance principle, it is interesting to note that a counter-example to Theorem \ref{CLT2} in the ergodic setup also can be written. Namely, let $d\ge 2$, write for $x\in\mathbb{Z}^d$, $z(x)=(x_2,\cdots, x_d)\in\mathbb{Z}^{d-1}$. Let $\{\varepsilon_z\}_{z\in\mathbb{Z}^{d-1}}$ be i.i.d random variables with support in $(0, 1/2)$ and set \begin{equation} \omega(x, e)=\left\{ \begin{array}{rl} \varepsilon_{z(x)}, & \text{if } e=\pm e_1\\ (1-2\varepsilon_{z(x)})/2(d-1), & \text{else }. \end{array} \right. \end{equation} It is easy to verify that $\{X_t^n\}_{t\ge 0}$ satisfies the quenched invariance principle, but that the limiting covariance may degenerate if the tail of $\varepsilon_z$ is heavy. \chapter{Einstein Relation for Random Walks in Balanced Random Environment} \label{ER chapter} In this chapter we will give the proof of the Einstein relation \eqref{Einstein relation} in the context of random walks in a balanced uniformly elliptic iid random environment. As mentioned in Section~\ref{IER}, our proof consists of proving Theorem~\ref{ER1} and Theorem~\ref{ER2}. We will prove Theorem \ref{ER1} in Section \ref{ERsec1}. In Section \ref{ERsecreg}, we will present our new construction of the regeneration times. Furthermore, we will show in Section \ref{ERsecmo} that these regeneration times have good moment properties. Section \ref{ERsecpro} is devoted to the proof of Theorem \ref{ER2}, using the regeneration times and arguments similar to \cite[pages 219-222]{GMP}. Throughout this chapter, we assume \textit{the environment $P$ is iid, balanced, and uniformly elliptic with ellipticity constant $\kappa>0$.} Recall that we have obtained in Section~\ref{SePeEn} an ergodic measure $Q$ for the process $\bar\omega(n)$. By the ergodic theorem, we get \[ \bm{D}=\big(2E_Q\omega(o,e_i)\delta_{ij}\big)_{1\le i,j\le d}, \] where $\bm D$ is the covariance matrix defined at the beginning of Section~\ref{IER}. \section{Proof of Theorem \ref{ER1}}\label{ERsec1} \begin{lemma}\label{ERl1} For any $t>0$ and any bounded continuous functional $F$ on $C([0,t],\mathbb{R}^d)$, \[ \mathop{\rm lim}_{\lambda\to 0}E_{\omega^\lambda} F(\lambda X_{s/\lambda^2};0\le s\le t) = EF(N_s+D_\ell s;0\le s\le t), \] where $(N_s)_{s\ge 0}$ is a $d$-dimensional Brownian motion with covariance matrix $\bm D$. \end{lemma} {\it Proof:}~ We first consider the Radon-Nikodym derivative of the measure $P_{\omega^\lambda}$ with respect to $P_\omega$. Put \[ G(t, \lambda)=G(t,\lambda;X_\cdot):=\log\prod_{j=1}^{\lceil t\rceil}[1+\lambda\ell\cdot(X_j-X_{j-1})]. \] Then \[ E_{\omega^\lambda} F(X_s: 0\le s\le t) = E_\omega F(X_s: 0\le s\le t)e^{G(t,\lambda)}. \] In particular, taking $F\equiv 1$, we have \begin{equation}\label{ERe0} E_\omega e^{G(t,\lambda)}=1 \end{equation} for any $\lambda\in (0,1)$ and $t>0$. Moreover, by the inequality $a-\frac{a^2}{2}\le \log (1+a)\le a-\frac{a^2}{2}+\frac{a^3}{3}$ for $a>0$, we get \begin{align}\label{ERe1} G(t,\lambda) &= \sum_{j=1}^{\lceil t \rceil}\log(1+\lambda\ell\cdot(X_j-X_{j-1}))\nonumber\\ &=\sum_{j=1}^{\lceil t \rceil} \left[\lambda\ell\cdot(X_j-X_{j-1})-\frac{\lambda^2\big(\ell\cdot(X_j-X_{j-1})\big)^2}{2}\right] +\lambda^2\lceil t\rceil H(\lambda)\nonumber\\ &=\lambda X_{\lceil t \rceil}\cdot\ell -\frac{\lambda^2}{2}\sum_{j=1}^{\lceil t \rceil}\big(\ell\cdot(X_j-X_{j-1})\big)^2 +\lambda^2\lceil t\rceil H(\lambda), \end{align} where the random variable $H(\lambda)=H(\lambda;X_\cdot)$ satisfies $0\le H\le \lambda/3$. Setting $h(\omega)=\sum_{i=1}^d\omega(o,e_i)\ell_i^2$, \[ \left( \sum_{j=1}^n \big(\ell\cdot(X_j-X_{j-1})\big)^2-2h(\omega_{X_{j-1}}) \right)_{n\ge 0} \] is a martingale sequence with bounded increments. Thus $P_\omega$-almost surely, \[ \mathop{\rm lim}_{n\to\infty}\frac{1}{n}\sum_{j=1}^n [\big(\ell\cdot(X_j-X_{j-1})\big)^2-2h(\theta^{X_{j-1}}\omega)]=0. \] Further, by the ergodic theorem, $P\otimes P_\omega$-almost surely, \begin{equation}\label{ERe2} \mathop{\rm lim}_{\lambda\to 0} \lambda^2\sum_{j=1}^{\lceil t/\lambda^2 \rceil} \big(\ell\cdot(X_j-X_{j-1})\big)^2 = \mathop{\rm lim}_{\lambda\to 0} \lambda^2\sum_{j=1}^{\lceil t/\lambda^2 \rceil} 2h(\theta^{X_{j-1}}\omega) = 2tE_Q h. \end{equation} We deduce from (\ref{ERe1}) and (\ref{ERe2}) that \[ e^{G(t/\lambda^2,\lambda)} =\exp[\lambda X_{t/\lambda^2}\cdot\ell-tE_Q h+O_{\lambda,X_\cdot}(1)], \] where $O_{\lambda,X_\cdot}(1)$ denotes a quantity that depends on $\lambda$ and $X_\cdot$, and $O_{\lambda,X_\cdot}(1)\to 0$ $P_\omega$-almost surely as $\lambda\to 0$. By Theorem~\ref{LaThm}, $(\lambda X_{s/\lambda^2})_{s\ge 0}$ converges weakly (under $P_\omega$) to $(N_s)_{s\ge 0}$. Hence for $P$-almost all $\omega$, \begin{equation}\label{ERe3} F(\lambda X_{s/\lambda^2};0\le s\le t)e^{G(t/\lambda^2,\lambda)} \end{equation} converges weakly (under $P_\omega$) to \[ F(N_s:0\le s\le t)\exp(N_t\cdot\ell-tE_Qh). \] Next, we will prove that for $P$-almost every $\omega$, this convergence is also in $L^1(P_\omega)$. It suffices to show that the class $(e^{G(t/\lambda^2,\lambda)})_{\lambda\in (0,1)}$ is uniformly integrable under $P_\omega$, $P$-a.s.. Indeed, for any $\gamma>1$, it follows from (\ref{ERe1}) and the estimate on $H(\lambda)$ that \begin{align*} &\gamma G(t/\lambda^2,\lambda)\\ &\le G(t/\lambda^2, \gamma\lambda) + \frac{(\gamma^2-\gamma)\lambda^2}{2} \sum_{j=1}^{\lceil t/\lambda^2 \rceil} \big(\ell\cdot(X_j-X_{j-1})\big)^2 +\gamma\lambda^2\lceil t/\lambda^2\rceil H(\lambda)\\ &< G(t/\lambda^2, \gamma\lambda) +\gamma^2 (t+1). \end{align*} Hence for $\gamma>1$ and all $\lambda\in (0,1)$, \[ E_\omega \exp(\gamma G(t/\lambda^2,\lambda)) \le e^{\gamma^2 (t+1)}E_\omega \exp(G(t/\lambda^2, \gamma\lambda)) \stackrel{by (\ref{ERe0})}{=}e^{\gamma^2 (t+1)}, \] which implies the uniform integrability of $(e^{G(t/\lambda^2,\lambda)})_{\lambda\in (0,1)}$. So the $L^1(P_\omega)$-convergence of (\ref{ERe3}) is proved and (for $P$-almost every $\omega$) we have \begin{align*} &\mathop{\rm lim}_{\lambda\to 0} E_{\omega^\lambda}F(\lambda X_{s/\lambda^2};0\le s\le t)\\ &=\mathop{\rm lim}_{\lambda\to 0} E_{\omega}F(\lambda X_{s/\lambda^2};0\le s\le t)e^{G(t/\lambda^2,\lambda)}\\ &=E \big[F(N_s:0\le s\le t)\exp(N_t\cdot\ell-t E_Qh)\big]. \end{align*} The lemma follows by noting that $tE_Qh=E(N_t\cdot\ell)^2/2$ and that, by Girsanov's formula, \[ E \big[F(N_s:0\le s\le t)\exp(N_t\cdot\ell-E(N_t\cdot\ell)^2/2)\big] = E F(N_s+D_\ell s;0\le s\le t). \]\qed \begin{lemma}\label{ERl2} For any $\lambda\in (0,1), t\ge 1/\lambda^2, p\ge 1$ and any balanced environment $\omega$, \[ E_{\omega^\lambda}\max_{0\le s\le t}|X_s|^p\le C_{p,d}(\lambda t)^p. \] Here we use $C_{p,d}$ to denote constants which depend only on $p$ and the dimension $d$, and which may differ from line to line. \end{lemma} {\it Proof:}~ Since the drift of $\omega^\lambda$ at $X_n, n\in\mathbb{N}$, is \begin{align*} E_{\omega^\lambda}(X_{n+1}-X_n|X_n) &=\sum_{|e|=1}\omega(X_n,e)(1+\lambda e\cdot\ell)e\\ &=\lambda\sum_{i=1}^d 2\omega(X_n,e_i)\ell_i e_i:=\lambda d_\omega(X_{n}), \end{align*} we get that \begin{equation}\label{ERe4} Y_n:=\lambda\sum_{i=1}^{n}d_\omega(X_{i-1})-X_n \end{equation} is a $P_{\omega^\lambda}$-martingale with bounded increments. By the Azuma-Hoeffding inequality, we get that for any $p\ge 1$, \[ E_{\omega^\lambda}\max_{1\le i\le n}|Y_i|^p \le C_{p,d}n^{p/2}. \] Hence \[ E_{\omega^\lambda}\max_{1\le i\le n}|X_i|^p \le 2^p(E_{\omega^\lambda}\max_{1\le i\le n}|Y_i|^p+\lambda^p n^p) \le C_{p,d}\lambda^p n^p \] for any $n\ge 1/\lambda^2$. The same inequality is true (with different $C_{p,d}$) if we replace $n\in\mathbb{N}$ with any $t\in\mathbb{R}$ such that $t\ge 1/\lambda^2$. \qed\\ \textit{Proof of Theorem \ref{ER1}:} Note that Lemma \ref{ERl1} implies that $\lambda X_{t/\lambda^2}$ (under the law $P_{\omega^\lambda}$) converges weakly to $N_t+D_\ell t$ as $\lambda\to 0$. When $t\ge 1$, the uniform integrability of $(\lambda X_{t/\lambda^2})_{\lambda\in (0,1)}$ under the corresponding measures $P_{\omega^\lambda}$, as shown in Lemma \ref{ERl2}, then yields that this convergence is also in $L^1$.\qed \section{Regenerations}\label{ERsecreg} \subsection{Auxiliary estimates} For the rest of this section, we assume that $\ell_1=\ell\cdot e_1>0$. Let \[ \lambda_1:=\big(\lceil(2\lambda\ell_1)^{-1}\rceil\big)^{-1}/2, \] so that $0.5/\lambda_1$ is an integer. Note that \[ \frac{1}{2\lambda\ell_1}\le \frac{1}{2\lambda_1}<\frac{1}{2\lambda\ell_1}+1. \] For any $n\in\mathbb{Z}, x\in\mathbb{Z}^d$, call \[ \mathcal{H}_n^x=\mathcal{H}_n^x(\lambda,\ell):= \{y\in\mathbb{Z}^d: (y-x)\cdot e_1=n/\lambda_1\} \] \textit{the $n$-th level} (with respect to $x$). Denote the hitting time of the $n$-th level by \[ T_n=T_n(X_\cdot):=\inf\{t\ge 0: (X_t-X_0)\cdot e_1=n/\lambda_1\}, n\in\mathbb{Z}. \] Also set \[ T_{\pm 0.5}:=\inf\{t\ge 0: (X_t-X_0)\cdot e_1=\pm 0.5/\lambda_1\}. \] Since $\ell_1>0$, the random walk is transient in the $e_1$ direction. Thus $(T_n)_{n\ge 0}$ are finite $P_{\omega^\lambda}$-almost surely. \begin{proposition}\label{ERprop5} For any $n,m\in\mathbb{Z}^+$ and any balanced environment $\omega$, \[ P_{\omega^\lambda}(T_n<T_{-m}) = \dfrac{1-q_\lambda^m}{1-q_\lambda^{m+n}}, \] where $q_\lambda:=(\frac{1-\lambda\ell_1}{1+\lambda\ell_1})^{1/\lambda_1}$. \end{proposition} {\it Proof:}~ Observe that the jumps of $(X_n\cdot e_1)_{n\ge 0}$ are lazy random walks on $\mathbb{Z}$, with the ratio of the probabilities of left-jump to right-jump equals $(1-\lambda\ell_1)/(1+\lambda\ell_1)$. Hence for $i,j\in\mathbb{Z}^+$, \[ P_{\omega^\lambda}(\tilde{T}_i<\tilde{T}_{-j}) =\frac{1-(\frac{1-\lambda\ell_1}{1+\lambda\ell_1})^j} {1-(\frac{1-\lambda\ell_1}{1+\lambda\ell_1})^{i+j}}, \] where $\tilde{T}_k:=\inf\{n\ge 0: (X_n-X_0)\cdot e_1=k\}, k\in\mathbb{Z}$. The proposition follows by noting that $T_n=\tilde{T}_{n/\lambda_1}$. \qed \begin{lemma}\label{ERl3} For all $\lambda\in(0,1), t>0, m\in\mathbb{N}$ and any balanced environment $\omega$ with ellipticity constant $\kappa\in(0,1/(2d))$, \[ P_{\omega^\lambda}(T_m\ge t/\lambda_1^2) \le 2 e^{-t\kappa^2/(2m)}. \] \end{lemma} {\it Proof:}~ First, note that if $Z$ is a real-valued random variable with zero mean and supported on $[-c,c]$, then for $\theta>0$, $Ee^{\theta Z}\le \exp{(\frac{1}{2}\theta^2c^2)}$. (By Jensen's inequality, $e^{\theta Z}\le \frac{c-Z}{2c}e^{\theta c}+\frac{c+Z}{2c}e^{-\theta c}$. Taking expectations on both sides gives the inequality.) Recall the definition of $Y_n$ in (\ref{ERe4}). Since $Y_n\cdot e_1$ is a $P_{\omega^\lambda}$-martingale with increments bounded by $2$, for $\theta>0$, \begin{align*} &E_{\omega^\lambda}(e^{\theta Y_{n+1}\cdot e_1}|X_i,i\le n)\\ &= e^{\theta Y_n\cdot e_1}E_{\omega^\lambda}[e^{\theta(Y_{n+1}-Y_n)\cdot e_1}||X_i,i\le n] \le e^{\theta Y_n\cdot e_1+2\theta^2}. \end{align*} Hence \[ \exp{(\theta Y_n\cdot e_1-2n\theta^2)} \] is a $P_{\omega^\lambda}$-supermartingale. By the optional stopping theorem and ellipticity, \begin{align*} 1 & \ge E_{\omega^\lambda} \exp[\theta Y_{T_m}\cdot e_1-2 T_m\theta^2]\\ & \ge E_{\omega^\lambda} \exp[\theta(2\lambda\ell_1\kappa T_m-X_{T_m}\cdot e_1)-2T_m\theta^2]. \end{align*} Letting $\theta=\kappa\lambda\ell_1/2$ in the above inequality and noting that $X_{T_m}\cdot e_1=m/\lambda_1$, we obtain \begin{align*} 1 &\ge E_{\omega^\lambda}\exp\big((\kappa\lambda\ell_1)^2T_m/2-\kappa\lambda\ell_1 m/(2\lambda_1)\big)\\ &\ge E_{\omega^\lambda}\exp(\kappa^2\lambda_1^2T_m/2-\kappa m), \end{align*} where we used $\lambda_1\le \lambda\ell_1\le 2\lambda_1$ in the second inequality. Hence by H\"{o}lder's inequality, \[ E_{\omega^\lambda}\exp(\kappa^2\lambda_1^2T_m/(2m)-\kappa)\le 1. \] Therefore, \[ P_{\omega^\lambda}(T_m\ge t/\lambda_1^2) \le e^{\kappa-\kappa^2 t/(2m)} <2 e^{-\kappa^2 t/(2m)}. \qed \] \begin{proposition}\label{ERprop2} There exists a constant $C_0=C_0(\kappa,d)>0$ such that \[ P_{\omega^\lambda}(\max_{0\le s\le T_1}|X_s|\ge C_0/\lambda_1)<0.5. \] \end{proposition} {\it Proof:}~ By Lemma \ref{ERl2} and Lemma \ref{ERl3}, for any $m\ge 1$, \begin{align*} &P_{\omega^\lambda}(\max_{0\le s\le T_1}|X_s|\ge m/\lambda_1)\\ &\le P_{\omega^\lambda}(T_1\ge \sqrt{m}/\lambda_1^2) +P_{\omega^\lambda}(\max_{0\le s\le\sqrt{m}/\lambda_1^2}|X_s|\ge m/\lambda_1)\\ &\le 2e^{-\sqrt{m}\kappa^2/2}+C/\sqrt{m}, \end{align*} which is less than $0.5$ if $m$ is large enough.\qed \begin{lemma}\label{ERl4} There exists a constant $c_1\in (0,1]$ such that for any $\lambda\in (0,1)$, $x\in\mathbb{Z}^d$ and balanced environment $\omega$, \begin{equation}\label{ERe5} P_{\omega^\lambda}^{x}(X_{T_1}=\cdot) \ge c_1 P_{\omega^\lambda}^{x+0.5e_1/\lambda_1}(X_{T_{0.5}}=\cdot|T_{0.5}<T_{-0.5}). \end{equation} \end{lemma} {\it Proof:}~ For any $x\in\mathbb{Z}^d$, let \[ \mathcal{H}_{0.5}^x :=\{y\in\mathbb{Z}^d: (y-x)\cdot e_1=0.5/\lambda_1\}. \] Fix $w\in\mathcal{H}_1^x$. Then the function \[ f(z):= P_{\omega^\lambda}^z(X_\cdot \text{ visits $\mathcal{H}_1^x$ for the first time at }w) \] satisfies \[ L_{\omega^\lambda}f(z)=0 \] for all $z\in\{y: (y-x)\cdot e_1<1/\lambda_1\}$. By the Harnack inequality for discrete harmonic functions (See Theorem \ref{ERharnack} in the Appendix. In this case $a=\omega^\lambda, R=0.5/\lambda_1$ and $b_0\le\lambda$), there exists a constant $C_2$ such that, for any $y, z\in \mathcal{H}_{0.5}^x$ with $|z-y|<0.5/\lambda_1$, \[ f(z)\ge C_2 f(y). \] Hence, for any $z\in \mathcal{H}_{0.5}^x$ such that $|z-(x+0.5e_1/\lambda_1)|<C_0/\lambda_1$, we have \begin{equation}\label{ERe27} f(z)\ge C_2^{2C_0} f(x+0.5e_1/\lambda_1). \end{equation} Therefore, \begin{align*} P_{\omega^\lambda}^x(X_{T_1}=w) &\ge \sum_{|y-x|<C_0/\lambda} P_{\omega^\lambda}^x(X_{T_{0.5}}=y)P_{\omega^\lambda}^y(X_{T_{0.5}}=w)\\ &\stackrel{\eqref{ERe27}}{\ge} C P_{\omega^\lambda}^x(|X_{T_{0.5}}-x|<C_0/\lambda_1)P_{\omega^\lambda}^{x+0.5e_1/\lambda_1} (X_{T_{0.5}}=w)\\ &\ge c_1 P_{\omega^\lambda}^{x+0.5e_1/\lambda_1} (X_{T_{0.5}}=w|T_{0.5}<T_{-0.5}) \end{align*} where in the last inequality we used the facts that (by Proposition \ref{ERprop2}) \[ P_{\omega^\lambda}^x(|X_{T_{0.5}}-x|<C_0/\lambda_1)>\frac{1}{2} \] and \[ P_{\omega^\lambda}^{x+0.5e_1/\lambda_1}(T_{0.5}<T_{-0.5})>\frac{1}{2}.\qed \] \subsection{Construction of the regeneration times} Let \[ \mu_{\omega^\lambda,1}^x(\cdot)=P_{\omega^\lambda}^{x+0.5e_1/\lambda_1} (X_{T_{0.5}}=\cdot|T_{0.5}<T_{-0.5}). \] Recall that $c_1$ is the constant in Lemma \ref{ERl4}. For any $\beta\in (0,c_1)$, we set \[ \mu_{\omega^\lambda,0}^x(\cdot)=\mu_{\omega^\lambda,0}^{x,\beta}(\cdot) := \big[P_{\omega^\lambda}^x(X_{T_1}=\cdot)-\beta\mu_{\omega^\lambda,1}^x(\cdot)\big]/(1-\beta). \] Then by (\ref{ERe5}), both $\mu_{\omega^\lambda,1}^x$ and $\mu_{\omega^\lambda,0}^x$ are probability measures on $\mathcal{H}^x_{0.5}$ and \[ P_{\omega^\lambda}^x(X_{T_1}=u)=\beta\mu_{\omega^\lambda,1}^x(u)+(1-\beta)\mu_{\omega^\lambda,0}^x(u). \] For any $\mathcal{O}\in\sigma(X_1,X_2,\ldots, X_{T_1}), x\in\mathbb{Z}^d$ and $i\in\{0,1\}$, put \begin{align}\label{ERe6} \nu_{\omega^\lambda,i}^x(\mathcal{O}) &= \nu_{\omega^\lambda,i}^{x,\beta}(\mathcal{O})\nonumber\\ &:= \sum_y \big[i\mu_{\omega^\lambda,1}^x(y)+(1-i)\mu_{\omega^\lambda,0}^x(y)\big] P_{\omega^\lambda}^x(\mathcal{O}|X_{T_1}=y). \end{align} Notice that under the environment measure $P$, \[ \nu_{\omega^\lambda,1}^x(X_{T_1}\in\cdot)=\mu_{\omega^\lambda,1}^x(\cdot) \] is independent of $\sigma(\omega_y:y\cdot e_1\le x\cdot e_1)$.\\ We will now define the regeneration times. We first sample a sequence $(\epsilon_i)_{i=1}^\infty\in\{0,1\}^\mathbb{N}$ of iid Bernoulli random variables according to the law $Q_\beta$ defined by \[ Q_\beta(\epsilon_i=1)=\beta \text{ and } Q_\beta(\epsilon_i=0)=1-\beta. \] Then, fixing $\epsilon:=(\epsilon_i)_{i=1}^\infty$, we will define a new law $P_{\omega^\lambda,\epsilon}$ on the paths as follows (see Figure \ref{ERfig0}). For $x\in\mathbb{Z}^d$, set \[P_{\omega^\lambda,\epsilon}^x(X_0=x)=1.\] Assume that the $P_{\omega^\lambda,\epsilon}^x$-law for finite paths of length$\le n$ is defined. For any path $(x_i)_{i=0}^{n+1}$ with $x_0=x$, define \begin{align*} &P_{\omega^\lambda,\epsilon}^{x} (X_{n+1}=x_{n+1},\ldots, X_{0}=x_0)\\ &:= P_{\omega,\epsilon}^{x}(X_I=x_I,\ldots, X_0=x_0) \nu_{\omega^\lambda,\epsilon_J}^{x_I}(X_{n+1-I}=x_{n+1},\ldots, X_1=x_{I+1}), \end{align*} where \[ J=J(x_0,\ldots,x_n):=\max\{j\ge 0: \mathcal{H}_{j}^{x_0}\cap\{x_i, 0\le i\le n\}\neq\emptyset\} \] is the highest level visited by $(x_i)_{i=0}^{n}$ and \[ I=I(x_0,\ldots,x_n):=\min\{0\le i\le n: x_i\in\mathcal{H}_J^{x_0}\} \] is the hitting time to the $J$-th level. By induction, the law $P_{\omega^\lambda,\epsilon}^x$ is well-defined for paths of all lengths. \begin{figure}[h] \centering \includegraphics[width=0.9\textwidth]{law} \caption{The law $\bar P_{\omega^\lambda,\epsilon}$ for the walks.} \label{ERfig0} \end{figure} Note that a path sampled by $P_{\omega^\lambda,\epsilon}^x$ is not a Markov chain, but the law of $X_\cdot$ under \[ \bar{P}_{\omega^\lambda}^x=\bar{P}_{\omega^\lambda,\beta}^{x}:=Q_\beta\otimes P_{\omega^\lambda,\epsilon}^x \] coincides with $P_{\omega^\lambda}^x$. That is, \[ \bar{P}_{\omega^\lambda}^x(X_\cdot\in\cdot) = P_{\omega^\lambda}^x(X_\cdot\in\cdot). \] Denote by $\bar{\mathbb P}_\lambda=\bar{\mathbb P}_{\lambda,\beta}:=P\otimes\bar{P}_{\omega^\lambda,\beta}$ the law of the triple $(\omega,\epsilon, X_\cdot)$. Expectations with respect to $\bar{P}_{\omega^\lambda}^x$ and $\bar{\mathbb P}_\lambda$ are denoted by $\bar{E}_{\omega^\lambda}^x$ and $\bar{\mathbb E}_\lambda (=\bar{\mathbb E}_{\lambda,\beta})$, respectively. Next, for a path $(X_n)_{n\ge 0}$ sampled according to $P_{\omega^\lambda,\epsilon}^o$, we will define the regeneration times. See Figure \ref{ERfig2} for an illustration. \begin{figure}[h] \centering \includegraphics[width=0.9\textwidth]{defreg} \caption{The definition of a regeneration time.} \label{ERfig2} \end{figure} To be specific, put $S_0=0, M_0=0$, and define inductively \begin{align*} &S_{k+1}=\inf\{T_{n+1}: n/\lambda_1\ge M_k \text{ and }\epsilon_n=1\},\\ &R_{k+1}=S_{k+1}+T_{-1}\circ\theta_{S_{k+1}},\\ &M_{k+1}=X_{S_{k+1}}\cdot e_1+N\circ \theta_{S_{k+1}}, \qquad k\ge 0. \end{align*} Here $\theta_n$ denotes the time shift of the path, i.e, $\theta_n X_\cdot=(X_{n+i})_{i=0}^\infty$, and \[ N:=\inf\{n/\lambda_1: n/\lambda_1>(X_i-X_0)\cdot e_1 \text{ for all }i\le T_{-1}\}. \] Set \begin{align*} &K:=\inf\{k\ge 1: S_k<\infty, R_k=\infty\},\\ &\tau_1:=S_K,\\ &\tau_{k+1}=\tau_k+\tau_1\circ\theta_{\tau_k}. \end{align*} We call $(\tau_k)_{k\ge 1}$ the ($\beta$-)\textit{regeneration times}. Intuitively, under $\bar{P}_{\omega^\lambda}^x$, whenever the walker visits a new level $\mathcal{H}_i, i\ge 0$, he flips a coin $\epsilon_i$. If $\epsilon_i=0$ (or $1$), he then walks following the law $\nu_{\omega^\lambda,0}$ (or $\nu_{\omega^\lambda,1}$) until he hits the $(i+1)$-th level. The regeneration time $\tau_1$ is defined to be the first time of visiting a new level $\mathcal{H}_k$ such that the outcome $\epsilon_{k-1}$ of the previous coin-tossing is ``$1$" and the path will never backtrack to the level $\mathcal{H}_{k-1}$ in the future. See Figure \ref{ERfig1}. \begin{figure}[h] \centering \includegraphics[width=0.8\textwidth]{regenerations} \caption{In this picture, $K=2, X_{\tau_1}=5/\lambda_1, M_1=4/\lambda_1$.} \label{ERfig1} \end{figure} \subsection{The renewal property of the regenerations} The regeneration times possess good renewal properties in the following sense: \begin{enumerate} \item Since the ratio of the probabilities of left-jump and right-jump of the lazy random walks $(X_n\cdot e_1)_{n\ge 0}$ (in $\mathbb{Z}$) is $(1-\lambda\ell_1)/(1+\lambda\ell_1)$, the law of $(X_{\tau_n}\cdot e_1)_{n\ge 1}$ does not depend on the environment $\omega$. (Indeed, if we only observe the chain $(X_n\cdot e_1)_{n\ge 0}$ at the times when it \textit{moves} and forget about its laziness, we get a random walk on $\mathbb{Z}$ with probabilities $(1-\lambda\ell_1)/2$ and $(1+\lambda\ell_1)/2$ of jumping to the left and to the right, respectively.) Furthermore, under $\bar{P}_{\omega^\lambda}$, the inter-regeneration distances $(e_1\cdot X_{\tau_1}\circ\theta_{\tau_n})_{n=1}^\infty$ in the direction $e_1$ are iid random variables which are independent of $X_{\tau_1}\cdot e_1$, and \[ \bar{P}_{\omega^\lambda}(e_1\cdot X_{\tau_1}\circ\theta_{\tau_n}\in\cdot) = \bar{P}_{\omega^\lambda}(X_{\tau_1}\cdot e_1\in\cdot|T_{-1}=\infty), n\ge 1. \] \item For $k\ge 0$, define \begin{align*} &\tilde{S}_{k+1}:=\inf\{T_n: n/\lambda\ge M_k \text{ and }\epsilon_n=1\},\\ &\tilde{\tau}_1:=\tilde{S}_K,\\ &\tilde{\tau}_{k+1}:=\tau_k+\tilde{\tau}_1\circ\theta_{\tau_k}. \end{align*} Note that for $k\ge 1$, \begin{align*} &S_{k}=\tilde S_{k}+T_1\circ\theta_{\tilde S_k},\\ &X_{\tau_k}\cdot e_1 =X_{\tilde{\tau}_k}\cdot e_1+1/\lambda_1. \end{align*} Conditioning on $X_{\tilde{\tau}_k}=x$, the law of $X_{\tau_k}$ is $\mu_{\omega^\lambda,1}^x$, which is independent (under the environment measure $P$) of $\sigma(\omega_y:y\cdot e_1\le x\cdot e_1)$. Moreover, after time $\tau_k$, the path will never visit $\{y:y\cdot e_1\le x\cdot e_1\}$. Thus the movement of the path after time $\tau_k$ is independent (under $\bar{\mathbb P}_\lambda$) of $(X_n)_{n\le \tilde{\tau}_k}$, and therefore, we expect \[ (\tilde{\tau}_1\circ\theta_{\tau_k})_{k\ge 1} \] to be iid random variables under $\bar{\mathbb P}_\lambda$. See Proposition \ref{ERprop1} for a rigorous proof. \item Although the inter-regeneration distances $(X_{\tau_1}\circ\theta_{\tau_k})_{k\ge 1}$ and $(\tilde{\tau}_1\circ\theta_{\tau_k})_{k\ge 1}$ are both iid sequences, the inter-regeneration times $(\tau_1\circ\theta_{\tau_k})_{k\ge 1}$ are not even independent. However, letting \[ \Delta_k:=T_1\circ\theta_{X_{\tilde{\tau}_k}}=\tau_k-\tilde{\tau}_k \text{ for }k\ge 1, \] we can show that for every $k\ge 1$, $\lambda_1^2 \Delta_k$ is bounded by a constant plus an exponential random variable. So $\Delta_k$ is much less than $\tau_1\circ\theta_{\tau_k}$, which is roughly $C/(\beta\lambda_1^2)$ (as will be shown in Proposition \ref{ERprop3}). In this sense, the inter-regeneration times $\tau_1\circ\theta_{\tau_k}$ are \textit{almost} iid if $\beta$ is sufficiently small. \end{enumerate} The rest of this subsection is devoted to the proof that $(\tilde{\tau}_1\circ\theta_{\tau_k})_{k\ge 1}$ are iid (Proposition \ref{ERprop1}) and that $\Delta_k$'s are dominated by iid random variables of sizes $1/\lambda^2$ (Proposition \ref{ERprop4}). We introduce the $\sigma$-field \[ \mathcal{G}_k := \sigma\big( \tilde{\tau}_k, (X_i)_{i\le\tilde{\tau}_k},(\omega_y)_{y\cdot e_1\le X_{\tilde{\tau}_k}\cdot e_1} \big). \] \begin{lemma}\label{ERl5} For any appropriate measurable sets $B_1, B_2$ and any event \[ B:=\{(X_i)_{i\ge 0}\in B_1, (\omega_y)_{y\cdot e_1>-1/\lambda_1}\in B_2\}, \] we have, for $k\ge 1$, \[ \bar{\mathbb P}_\lambda(B\circ\bar{\theta}_{\tau_k}|\mathcal{G}_k) = \frac{E_P\big[ \sum_y \mu_{\omega^\lambda,1}(y) \bar{P}_{\omega^\lambda}^y(B\cap\{T_{-1}=\infty\})\big]} { E_P\big[ \sum_y \mu_{\omega^\lambda,1}(y) \bar{P}_{\omega^\lambda}^y(T_{-1}=\infty)\big] }. \] Here $\bar{\theta}_n$ is the shift defined by \[ B\circ\bar{\theta}_n = \{(X_i)_{i\ge n}\in B_1, (\omega_y)_{(y-X_n)\cdot e_1>-1/\lambda_1}\in B_2\}. \] \end{lemma} {\it Proof:}~ For simplicity, let us consider the case $k=1$. We use $\theta^n$ to denote the shift of the $\epsilon$-coins, i.e., $\theta^n \epsilon_\cdot=(\epsilon_i)_{i\ge n}$. For any $A\in\mathcal{G}_1$, \begin{align*} &\bar{\mathbb P}_\lambda(B\circ\bar{\theta}_{\tau_1}\cap A)\\ &= E_{P\otimes Q_\beta}\big[ \sum_{k\ge 1,x}P_{\omega^\lambda,\epsilon} (A\cap\{\tilde{S}_k<\infty,R_k=\infty, X_{\tilde{S}_k}=x\}\cap B\circ\bar{\theta}_{S_k}) \big]\\ &= E_{P\otimes Q_\beta}\big[ \sum_{k\ge 1,x,y}P_{\omega^\lambda,\epsilon} (A\cap\{\tilde{S}_k<\infty,X_{\tilde{S}_k}=x\}) \nu_{\omega^\lambda,1}^x(X_{T_1}=x+y)\\ &\qquad\qquad\qquad\qquad\qquad\times P_{\omega^\lambda,\theta^{k+1}\epsilon}^{x+y}(B\cap\{T_{-1}=\infty\})\big]. \end{align*} Note that in the last equality, \[P_{\omega^\lambda,\epsilon} (A\cap\{\tilde{S}_k<\infty,X_{\tilde{S}_k}=x\})\] is $\sigma\big((\epsilon_i)_{i\le k},(\omega_z)_{(z-x)\cdot e_1\le 0}\big)$- measurable, whereas \[\nu_{\omega^\lambda,1}^x(X_{T_1}=x+y)P_{\omega^\lambda,\theta^{k+1}\epsilon}^{x+y}(B\cap\{T_{-1}=\infty\})\] is $\sigma\big((\epsilon_i)_{i\ge k+1}, (\omega_z)_{(z-x)\cdot e_1>0}\big)$- measurable for $y\in\mathcal{H}_1^x$. Hence they are independent under $P\otimes Q_\beta$ and we have \begin{align}\label{ERe7} &\bar{\mathbb P}_\lambda(B\circ\bar{\theta}_{\tau_1}\cap A)\\ &= \sum_{k\ge 1}\bar{\mathbb P}_\lambda (A\cap\{\tilde{S}_k<\infty\}) E_P\big[ \sum_y \nu_{\omega^\lambda,1}(X_{T_1}=y) \bar{P}_{\omega^\lambda}^y(B\cap\{T_{-1}=\infty\})\big].\nonumber \end{align} Substituting $B$ with the set of all events, we get \begin{equation}\label{ERe8} \bar{\mathbb P}_\lambda(A)= \sum_{k\ge 1}\bar{\mathbb P}_\lambda (A\cap\{\tilde{S}_k<\infty\}) E_P\big[ \sum_y \mu_{\omega^\lambda,1}(y) \bar{P}_{\omega^\lambda}^y(T_{-1}=\infty)\big]. \end{equation} (\ref{ERe7}) and (\ref{ERe8}) yield that \[ \bar{\mathbb P}_\lambda(B\circ\bar{\theta}_{\tau_1}|A) = \frac{E_P\big[ \sum_y \mu_{\omega^\lambda,1}(y) \bar{P}_{\omega^\lambda}^y(B\cap\{T_{-1}=\infty\})\big]} { E_P\big[ \sum_y \mu_{\omega^\lambda,1}(y) \bar{P}_{\omega^\lambda}^y(T_{-1}=\infty)\big] } .\] The lemma is proved for the case $k=1$. The general case $k>1$ follows by induction. (The reasoning for the induction step is the same, although the notation becomes more cumbersome.)\qed The following proposition is an immediate consequence of the lemma. \begin{proposition}\label{ERprop1} Under $\bar{\mathbb P}_\lambda$, $\tilde{\tau}_1,\tilde{\tau}_1\circ\theta_{\tau_1},\ldots, \tilde{\tau}_1\circ\theta_{\tau_k},\ldots$ are independent random variables. Furthermore, $(\tilde{\tau}_1\circ\theta_{\tau_k})_{k\ge 1}$ are iid with law \[ \bar{\mathbb P}_\lambda(\tilde{\tau}_1\circ\theta_{\tau_k}\in\cdot) = \frac{ E_P\big[ \sum_y \mu_{\omega^\lambda,1}(y) \bar{P}_{\omega^\lambda}^y(\tilde{\tau}_1\in\cdot,T_{-1}=\infty)\big] } { E_P\big[ \sum_y \mu_{\omega^\lambda,1}(y) \bar{P}_{\omega^\lambda}^y(T_{-1}=\infty)\big] }. \] \end{proposition} Note that the inter-regeneration times $(\tau_1\circ\theta_{\tau_k})_{k\ge 1}$ are not independent. However, the differences between $\tau_1\circ\theta_{\tau_k}$ and $\tilde{\tau}_1\circ\theta_{\tau_k}, k\ge 1$ are controlled by iid exponential random variables. For any $x\in\mathbb{Z}^d, t\ge 0$ , \begin{align*} &\nu_{\omega^\lambda,1}^x(\lambda_1^2T_1\ge t)\\ &\stackrel{\eqref{ERe6}}{=} \sum_y \mu_{\omega^\lambda,1}^x(y)P_{\omega^\lambda}^x(\lambda_1^2T_1\ge t|X_{T_1}=y)\\ &\stackrel{\eqref{ERe5}}{\le} c_1^{-1} \sum_y P_{\omega^\lambda}^x(X_{T_1}=y)P_{\omega^\lambda}^x(\lambda_1^2T_1\ge t|X_{T_1}=y)\\ &= c_1^{-1}P_{\omega^\lambda}^x(\lambda_1^2T_1\ge t) \stackrel{\text{Lemma }\ref{ERl3}}{\le} 2c_1^{-1}e^{-t\kappa^2/2}. \end{align*} Hence for $k\ge 1$, \[ P_{\omega^\lambda,\epsilon}(\lambda_1^2\Delta_k\ge t|X_i,i\le \tilde{\tau}_k)= \nu_{\omega^\lambda,1}^{X_{\tilde{\tau}_k}}(\lambda_1^2T_1\ge t) \le 2c_1^{-1}e^{-t\kappa^2/2}, \] which implies that $\lambda_1^2\Delta_k$ is stochastically dominated by an exponential random variable (with rate $\kappa^2/2$) plus a constant $c_2:=2\kappa^{-2}\log(2/c_1)$. Thus we conclude: \begin{proposition}\label{ERprop4} Enlarging the probability space if necessary, one can couple $(\Delta_k)_{k\ge 1}$ with an iid sequence $(\xi_k)_{k\ge 1}$ such that each $\xi_k$ is the sum of $c_2$ and an exponential random variable with rate $\kappa^2/2$, and that \[ \lambda_1^2\Delta_k\le \xi_k, \text{ for all }k\ge 1. \] Therefore, for any $n\ge 1$, \begin{equation}\label{ERe9} \tilde{\tau}_1+\sum_{i=1}^{n-1}\tilde{\tau}_1\circ\theta_{\tau_i} \le \tau_n \le \tilde{\tau}_1+\sum_{i=1}^{n-1}\tilde{\tau}_1\circ\theta_{\tau_i}+\sum_{i=1}^n\xi_i/\lambda_1^2. \end{equation} \end{proposition} \section{Moment estimates}\label{ERsecmo} Throughout this section, we assume that \[\ell\cdot e_1>0.\] Set $\tau_0=0$. We will show that the typical values of $e_1\cdot X_1\circ\theta_{\tau_k}$ and $\tau\circ\theta_{\tau_k}$, $k\ge 0$ are $C/(\beta\lambda)$ and $C/(\beta\lambda^2)$, respectively. \begin{theorem}\label{ERregdist} Let $\omega$ be an elliptic and balanced environment. If $\lambda>0$ and $\beta>0$ are small enough, then \[ \bar{E}_{\omega^\lambda}\exp(\beta\lambda_1 X_{\tau_1}\cdot e_1/2)<12. \] \end{theorem} {\it Proof:}~ For $0\le k\le K-1$, set \[ L_{k+1}=\inf\{n\ge \lambda_1 M_k: \epsilon_n=1\}-\lambda_1 M_k+1. \] Then $L_1$ is the number of coins tossed to get the first `$1$' and \[ X_{S_1}\cdot e_1=L_1/\lambda_1. \] Moreover, for $1\le k\le K-1$, let \[ N_k=N\circ\theta_{S_k}. \] Then \[ (X_{S_{k+1}}-X_{S_k})\cdot e_1=N_{k}+L_{k+1}/\lambda_1, \quad k\ge 1. \] So \begin{equation}\label{ERe10} X_{\tau_1}\cdot e_1=\sum_{i=1}^K L_i/\lambda_1+\sum_{i=1}^{K-1}N_i. \end{equation} First, we will compute the exponential moment of $L_i, i\le K$. Since $(L_i)_{i\ge 1}$ depends only on the coins $(\epsilon_i)_{i\ge 0}$, it is easily seen that they are iid geometric random variables with parameter $\beta$. Hence for $i\ge 1$ (noting that $(1-\beta)e^{\beta/2}< e^{-\beta/2}<1$), \[ \bar{E}_{\omega^\lambda} [e^{\beta L_i/2}] = \sum_{n=0}^\infty e^{\beta(n+1)/2}(1-\beta)^n\beta = \dfrac{\beta e^{\beta/2}}{1-(1-\beta)e^{\beta/2}}. \] If $\beta>0$ is small enough, we have \begin{equation}\label{ERe11} \bar{E}_{\omega^\lambda} [e^{\beta L_i/2}]<3. \end{equation} Next, we will compute the exponential moment of $N_i, i\le K-1$. By Proposition \ref{ERprop5}, putting \[ p_\lambda:=\bar{P}_{\omega^\lambda}(T_{-1}=\infty) =1-q_\lambda, \] we have \begin{align*} &\bar{P}_{\omega^\lambda}(N=(n+1)/\lambda_1)\\ &= \bar{P}_{\omega^\lambda}(T_n<T_{-1}<T_{n+1})\\ &= \bar{P}_{\omega^\lambda}(T_n<T_{-1})- \bar{P}_{\omega^\lambda}(T_{n+1}<T_{-1})\\ &= \frac{p_\lambda}{1-q_\lambda^{n+1}}-\frac{p_\lambda}{1-q_\lambda^{n+2}} =\dfrac{q_\lambda^{n+1}p_\lambda^2}{(1-q_\lambda^{n+1})(1-q_\lambda^{n+2})}, \quad n\ge 0. \end{align*} Observe that conditioning on $K$, $(N_i)_{1\le i<K}$ are iid under $\bar{P}_{\omega^\lambda}$. Hence \begin{align*} \bar{P}_{\omega^\lambda}(N_i=(n+1)/\lambda_1|K>i) &=\bar{P}_{\omega^\lambda}(N=(n+1)/\lambda_1|T_{-1}<\infty)\\ &= \dfrac{q_\lambda^{n}p_\lambda^2}{(1-q_\lambda^{n+1})(1-q_\lambda^{n+2})}\le q_\lambda^n, \end{align*} and \[ \bar{E}_{\omega^\lambda} [e^{\beta\lambda_1 N_i/2}|K>i] \le \dfrac{e^{\beta/2}}{1-e^{\beta/2}q_\lambda}. \] Noting that $\mathop{\rm lim}_{\lambda\to 0}q_\lambda=e^{-2}$, we can take both $\lambda$ and $\beta$ to be small enough such that \begin{equation}\label{ERe12} \bar{E}_{\omega^\lambda} [e^{\beta\lambda_1 N_i/2}|K>i] <\frac{1}{4q_\lambda}. \end{equation} Finally, note that, under $\bar{P}_{\omega^\lambda}=Q_\beta\otimes P_{\omega^\lambda,\epsilon}^o$, $K$ is a geometric random variable with success parameter $p_\lambda$, and $(L_i)_{1\le i\le K}$ and $(N_i)_{1\le i\le K}$ are iid sequences when conditioned on $K$. Therefore, by (\ref{ERe10}), (\ref{ERe11}) and (\ref{ERe12}), \[ \bar{E}_{\omega^\lambda}\exp(\beta\lambda_1 X_{\tau_1}\cdot e_1/2) \le \bar{E}_{\omega^\lambda} \frac{3^K}{(4q_\lambda)^{K-1}} =\sum_{n=0}^\infty \frac{3^{n+1}}{(4q_\lambda)^n}q_\lambda^np_\lambda <12 \] if both $\beta,\lambda>0$ are small enough. \qed \begin{corollary}\label{ERcor1} For $t\ge 1$ and small enough $\lambda, \beta>0$, \[ \bar{P}_{\omega^\lambda}(\beta\lambda_1^2\tau_1\ge t) \le 14\exp(-\kappa^2\sqrt{t}/4). \] \end{corollary} {\it Proof:}~ By Lemma \ref{ERl3} and Theorem \ref{ERregdist}, \begin{align*} &\bar{P}_{\omega^\lambda}(\beta\lambda_1^2\tau_1\ge t)\\ &\le \bar{P}_{\omega^\lambda}(\beta\lambda_1^2T_{\lceil\sqrt{t}/\beta\rceil}\ge t) +\bar{P}_{\omega^\lambda}(T_{\lceil\sqrt{t}/\beta\rceil}<\tau_1)\\ &\le 2\exp(-\frac{\kappa^2t/\beta}{2(\sqrt{t}/\beta+1)}) +\bar{P}_{\omega^\lambda}(\lceil\sqrt{t}/\beta\rceil/\lambda_1<X_{\tau_1}\cdot e_1)\\ &\le 2e^{-\kappa^2\sqrt{t}/4}+12e^{-\sqrt{t}/2} \le 14 e^{-\kappa^2\sqrt{t}/4}. \qed \end{align*} It follows from Corollary \ref{ERcor1} and Lemma \ref{ERl5} (and noting that $P_{\omega^\lambda}(T_{-1}=\infty)=p_\lambda>1/2$) that, for $k\ge 1$, \begin{equation}\label{ERe13} \bar{\mathbb P}_\lambda (\beta\lambda_1^2\tau_1\circ\theta_{\tau_k}\ge t) \le 28\exp(-\kappa^2\sqrt{t}/4). \end{equation} Hence by Theorem \ref{ERregdist}, Corollary \ref{ERcor1} and \eqref{ERe13}, we conclude that, for any $p\ge 1, k\ge 0$, there exists a constant $C(p)<\infty$ such that \begin{align} &\bar{\mathbb E}_\lambda (\beta\lambda_1^2\tau_1\circ\theta_{\tau_k})^p <C(p),\label{ERe14}\\ &\bar{\mathbb E}_\lambda (\beta\lambda_1 X_{\tau_1}\circ\theta_{\tau_k})^p <C(p).\label{ERe15} \end{align} Moreover, since $\bar{\mathbb P}_\lambda$-almost surely, \[ v_\lambda\cdot e_1=\mathop{\rm lim}_{n\to\infty}\frac{X_{\tau_n}\cdot e_1}{\tau_n}, \] by \eqref{ERe9} and the law of large numbers, we have \begin{equation}\label{ERe16} L^{\beta,\lambda}:=\dfrac{\bar{\mathbb E}_\lambda[e_1\cdot X_{\tau_1}\circ\theta_{\tau_1}]} {\bar{\mathbb E}_\lambda[\tilde{\tau}_1\circ\theta_{\tau_1}]+E\xi_1/\lambda_1^2} \le v_\lambda\cdot e_1 \le \dfrac{\bar{\mathbb E}_\lambda[e_1\cdot X_{\tau_1}\circ\theta_{\tau_1}]} {\bar{\mathbb E}_\lambda[\tilde{\tau}_1\circ\theta_{\tau_1}]}=:R^{\beta,\lambda}. \end{equation} \begin{proposition}\label{ERprop3} When $\lambda, \beta>0$ are small enough, \begin{equation}\label{ERe17} \bar{\mathbb E}_\lambda[\tilde{\tau}_1\circ\theta_{\tau_1}]\ge \frac{C}{\beta\lambda_1^2}. \end{equation} \end{proposition} {\it Proof:}~ By the definition of $L_i, i\ge 1$, we get \begin{equation}\label{ERe18} \bar{\mathbb E}_\lambda [e_1\cdot X_{\tau_1}\circ\theta_{\tau_1}] \ge \bar{\mathbb E}_\lambda L_1/\lambda_1 \ge \frac{1}{\beta\lambda_1}. \end{equation} On the other hand, Lemma \ref{ERl2} implies that \[ |v_\lambda|\le C\lambda \text{ for all }\lambda\in(0,1). \] This, together with \eqref{ERe16} and \eqref{ERe18}, yields \[ \bar{\mathbb E}_\lambda[\tilde{\tau}_1\circ\theta_{\tau_1}]+E\xi_1/\lambda_1^2 \ge \dfrac{C}{\beta\lambda_1^2}. \] Recalling (see Proposition \ref{ERprop4}) that $E\xi_1$ is an exponential random variable with rate $\kappa^2/2$, (\ref{ERe17}) then follows by taking $\beta$ sufficiently small.\qed Note that, by \eqref{ERe16} and \ref{ERe17}, \begin{equation}\label{ERe19} R^{\beta,\lambda}\le (1+C\beta)L^{\beta,\lambda}\le C\lambda. \end{equation} \section{Proof of the Einstein relation}\label{ERsecpro} \begin{lemma}\label{ERl6} Assume $\ell\cdot e_1>0$. Then when $\beta>0$ and $\lambda>0$ are small enough, there exists a constant $C$ such that \[ \left| \dfrac{\bar{\mathbb E}_\lambda X_{\tau_n}\cdot e_1}{\lambda\bar{\mathbb E}_\lambda\tau_n} -\frac{v_\lambda\cdot e_1}{\lambda} \right|\le C\beta+\frac{C}{n} \quad \text{ for all }n\ge 2. \] \end{lemma} {\it Proof:}~ For $n\ge 2$, since \[ \dfrac{\bar{\mathbb E}_\lambda X_{\tau_n}\cdot e_1}{\bar{\mathbb E}_\lambda\tau_n} \ge \dfrac{(n-1)\bar{\mathbb E}_\lambda[e_1\cdot X_{\tau_1}\circ\theta_{\tau_1}]} {\bar{\mathbb E}_\lambda\tau_1+(n-1)\big(\bar{\mathbb E}_\lambda[\tilde{\tau}_1\circ\theta_{\tau_1}]+E\xi_1/\lambda_1^2\big)}, \] and \[ \dfrac{\bar{\mathbb E}_\lambda X_{\tau_n}\cdot e_1}{\bar{\mathbb E}_\lambda\tau_n} \le \dfrac{\bar{\mathbb E}_\lambda X_{\tau_1}\cdot e_1+(n-1)\bar{\mathbb E}_\lambda[e_1\cdot X_{\tau_1}\circ\theta_{\tau_1}]} {(n-1)\bar{\mathbb E}_\lambda[\tilde{\tau}_1\circ\theta_{\tau_1}]}, \] by the moment bounds \eqref{ERe14}, \eqref{ERe15}, \eqref{ERe17} and \eqref{ERe18}, we have (for small $\beta$ and $\lambda$) \[ \frac{L^{\beta,\lambda}/\lambda}{C/(n-1)+1} \le \dfrac{\bar{\mathbb E}_\lambda X_{\tau_n}\cdot e_1}{\lambda\bar{\mathbb E}_\lambda\tau_n} \le \frac{C}{n-1}+\frac{R^{\beta,\lambda}}{\lambda}. \] Hence when $\beta>0$ and $\lambda>0$ are small enough and $n\ge 2$, by \eqref{ERe16}, \[ \left|\dfrac{\bar{\mathbb E}_\lambda X_{\tau_n}\cdot e_1}{\lambda\bar{\mathbb E}_\lambda\tau_n} -\frac{v_\lambda\cdot e_1}{\lambda}\right| \le \frac{C}{n-1}+\frac{R^{\beta,\lambda}}{\lambda}-\frac{L^{\beta,\lambda}/\lambda}{C/(n-1)+1} \stackrel{(\ref{ERe19})}{\le} C\beta+\frac{C}{n-1}. \] The lemma is proved.\qed \begin{lemma}\label{ERthm3} Assume $\ell\cdot e_1>0$. Let $\alpha_n=\alpha_n(\beta,\lambda):=\bar{\mathbb E}_\lambda\tau_n$. Then when $\beta>0$ and $\lambda>0$ are small enough, \[ \left| \dfrac{\bar{\mathbb E}_\lambda X_{\tau_n}\cdot e_1}{\lambda\alpha_n} -\dfrac{\bar{\mathbb E}_\lambda X_{\alpha_n}\cdot e_1}{\lambda\alpha_n} \right|\le \frac{C}{n^{1/4}}\quad\text{ for all }n\in \mathbb{N}. \] \end{lemma} Note that, by \eqref{ERe14} and \eqref{ERe17}, \begin{equation}\label{ERe20} \frac{Cn}{\beta\lambda^2} \le \alpha_n \le \frac{C(1)n}{\beta\lambda^2}. \end{equation} {\it Proof:}~ Assume that both $\lambda$ and $\beta$ are sufficiently small. First, for any $\rho\in(0,1)$, \begin{align}\label{ERe21} &\bar{\mathbb E}_\lambda [|X_{\alpha_n}-X_{\tau_n}|1_{|\tau_n-\alpha_n|\le \rho\alpha_n}]\\ &\le \bar{\mathbb E}_\lambda \big[\max_{(1-\rho)\alpha_n\le s\le (1+\rho)\alpha_n}|X_s-X_{\alpha_n}|\big]\stackrel{\text{Lemma \ref{ERl2}}}{\le} C\rho\lambda\alpha_n.\nonumber \end{align} Second, \begin{align}\label{ERe22} &\bar{\mathbb E}_\lambda [|(X_{\alpha_n}-X_{\tau_n})\cdot e_1|1_{|\tau_n-\alpha_n|> \rho\alpha_n}]\nonumber\\ &\le \sqrt{\bar{\mathbb E}_\lambda[|(X_{\alpha_n}-X_{\tau_n})\cdot e_1|^2] \bar{\mathbb P}_\lambda(|\tau_n-\alpha_n|> \rho\alpha_n)}\nonumber\\ &\stackrel{\text{Lemma \ref{ERl2}, \eqref{ERe15}}}{\le } Cn(\beta\lambda)^{-1}\sqrt{\bar{\mathbb P}_\lambda(|\tau_n-\alpha_n|> \rho\alpha_n)}. \end{align} Furthermore, we can show that \begin{equation}\label{ERe23} \bar{\mathbb P}_\lambda(|\tau_n-\alpha_n|> \rho\alpha_n)\le C/(n\rho^2). \end{equation} Indeed, put \[ A_n:=\tilde{\tau}_1+\sum_{i=1}^{n-1}\tilde{\tau}_1\circ\theta_{\tau_i} \] and $B_n:=A_n+\sum_{i=1}^n\xi_i/\lambda_1^2$. Then by (\ref{ERe9}), we have $A_n\le \tau_n\le B_n$. Thus \[ A_n-\bar{\mathbb E}_\lambda A_n-Cn/\lambda^2 \le \tau_n-\alpha_n \le B_n-\bar{\mathbb E}_\lambda B_n+Cn/\lambda^2. \] Hence, by \eqref{ERe20} and by taking $\beta>0$ small enough, we get \begin{align*} \bar{\mathbb P}_\lambda(\tau_n-\alpha_n> \rho\alpha_n) &\le \bar{\mathbb P}_\lambda(B_n-\bar{\mathbb E}_\lambda B_n\ge \rho\alpha_n/2)\\ &\le \frac{\mathop{\rm Var} B_n}{(\rho\alpha_n/2)^2}, \end{align*} and \begin{align*} \bar{\mathbb P}_\lambda(\tau_n-\alpha_n<-\rho\alpha_n) &\le \bar{\mathbb P}_\lambda(A_n-\bar{\mathbb E}_\lambda A_n\le -\rho\alpha_n/2)\\ &\le \frac{\mathop{\rm Var} A_n}{(\rho\alpha_n/2)^2}. \end{align*} Since (recalling Proposition \ref{ERprop1}) \[ \mathop{\rm Var} A_n=\mathop{\rm Var} \tilde{\tau}_1+(n-1)\mathop{\rm Var}\tilde{\tau}_1\circ\theta_{\tau_1} \stackrel{(\ref{ERe14})}{\le } Cn(\beta\lambda^2)^{-2} \] and \[ \mathop{\rm Var} B_n \le 2\big(\mathop{\rm Var} A_n+\mathop{\rm Var}(\sum_{i=1}^n\xi/\lambda_1^2)\big) = 2\mathop{\rm Var} A_n+Cn/\lambda_1^4 \le Cn(\beta\lambda^2)^{-2}, \] we conclude that \[ \bar{\mathbb P}_\lambda(|\tau_n-\alpha_n|> \rho\alpha_n) \le \frac{Cn(\beta\lambda^2)^{-2}}{(\rho\alpha_n/2)^2} \le C/(n\rho^2). \] This completes the proof of \eqref{ERe23}. Finally, combining (\ref{ERe21}), (\ref{ERe22}) and (\ref{ERe23}), we obtain \[ \left| \dfrac{\bar{\mathbb E}_\lambda X_{\tau_n}\cdot e_1}{\lambda\alpha_n} -\dfrac{\bar{\mathbb E}_\lambda X_{\alpha_n}\cdot e_1}{\lambda\alpha_n} \right| \le C\rho+\frac{C}{\rho\sqrt{n}}. \] The lemma follows by taking $\rho=\frac{1}{n^{1/4}}$.\qed\\ \noindent{\it Proof of Theorem \ref{ER2}:}\\ First, we will show that when $\lambda\in(0,1)$ is small enough, for any $t\ge 1$, \begin{equation}\label{ERe26} \left| \dfrac{\bar{\mathbb E}_\lambda X_{t/\lambda^2}\cdot e_1}{t/\lambda} -\frac{v_\lambda\cdot e_1}{\lambda} \right| \le \frac{C}{t^{1/5}}. \end{equation} Note that if $\ell\cdot e_1=0$, then $(X_n\cdot e_1)_{n=0}^\infty$ is a martingale and $\bar{\mathbb E}_\lambda X_n\cdot e_1=v_\lambda\cdot e_1=0$ for all $n$. Hence we only consider the non-trivial case $\ell\cdot e_1\neq 0$. Without loss of generality, assume $\ell\cdot e_1>0$. By Lemma \ref{ERl2}, the left side of \eqref{ERe26} is uniformly bounded for all $t\ge 1$ and $\lambda\in (0,1)$. So it suffices to prove \eqref{ERe26} for all sufficiently large $t>0$ and sufficiently small $\lambda>0$. When $t>0$ is sufficiently large and $\lambda>0$ is small enough, we let \begin{equation}\label{ERe24} \beta=\beta(t)=t^{-1/5} \end{equation} and set $n=n(t,\lambda)$ be the integer that satisfies \[ \alpha_n\le \frac{t}{\lambda^2}<\alpha_{n+1}. \] By \eqref{ERe20}, the existence of $n(t,\lambda)$ is guaranteed. Moreover, \begin{equation}\label{ERe25} n\ge Ct\beta= Ct^{-4/5}. \end{equation} Since \begin{align*} &\left| \dfrac{\bar{\mathbb E}_\lambda X_{\alpha_n}\cdot e_1}{\lambda\alpha_n}-\dfrac{\bar{\mathbb E}_\lambda X_{t/\lambda^2}\cdot e_1}{t/\lambda} \right|\\ &\le \frac{1}{\lambda\alpha_n}\bar{\mathbb E}_\lambda|X_{\alpha_n}-X_{t/\lambda^2}|+ \bar{\mathbb E}_\lambda|X_{t/\lambda}|(\frac{1}{\lambda\alpha_n}-\frac{1}{t})\\ &\le \frac{1}{\lambda\alpha_n} \bar{\mathbb E}_\lambda \big[\max_{\alpha_n\le s<\alpha_{n+1}}|X_{\alpha_n}-X_s|\big]+ \bar{\mathbb E}_\lambda[\max_{0\le s<\alpha_{n+1}}|X_s|] \frac{\lambda \bar{\mathbb E}_\lambda[\tau_1\circ\theta_{\tau_n}]}{(\lambda\alpha_n)^2}, \end{align*} by Lemma \ref{ERl2}, \eqref{ERe14} and \eqref{ERe20}, we obtain \begin{equation*} \left| \dfrac{\bar{\mathbb E}_\lambda X_{\alpha_n}\cdot e_1}{\lambda\alpha_n}- \dfrac{\bar{\mathbb E}_\lambda X_{t/\lambda^2}\cdot e_1}{t/\lambda} \right| \le \frac{C}{n}. \end{equation*} Combining Lemma \ref{ERl6}, Lemma \ref{ERthm3} and the above inequality, we conclude that if $t$ is sufficiently large and $\lambda>0$ is sufficiently small, then \[ \left| \dfrac{\bar{\mathbb E}_\lambda X_{t/\lambda^2}\cdot e_1}{t/\lambda} -\frac{v_\lambda\cdot e_1}{\lambda} \right|\le C\beta+\frac{C}{n^{1/4}} \le \frac{C}{t^{1/5}}. \] Here we used \eqref{ERe24} and \eqref{ERe25} in the last inequality. \eqref{ERe26} is proved. The same equality for the remaining directions $e_2,e_3,\ldots,e_d$ can be obtained using the same argument. Our proof of Theorem \ref{ER2} is complete. \qed \chapter{Limiting Velocity in Mixing Random Environment} \label{LV chapter} This chapter is devoted to the proof of Theorem~\ref{LVthm2}. The organization of the proof is as follows. In Section \ref{seccomb}, we prove a refined version of \cite[Lemma 3]{Ze}. With this combinatorial result, we will prove the CLLN \eqref{ICLLN} in Section \ref{seclln}, using coupling arguments. In Section \ref{sechke}, using coupling, we obtain heat kernel estimates, which is later used in Section \ref{secunique} to show the uniqueness of the non-zero limiting velocity. Throughout this chapter, we assume that \textit{the environment is uniformly elliptic with ellipticity constant $\kappa$ and satisfies $(G)$}. We use $c, C$ to denote finite positive constants that depend only on the dimension $d$ and the environment measure $P$ (and implicitly, on the parameters $\kappa,r$ and $\gamma$ of the environment). They may differ from line to line. We denote by $c_1,c_2,\ldots$ positive constants which are fixed throughout, and which depend only on $d$ and the measure $P$. Let $\{e_1,\ldots,e_d\}$ be the natural basis of $\mathbb{Z}^d$. \section{A combinatorial lemma and its consequences}\label{seccomb} In this section we consider the case that $\mathbb{P}(\varlimsup_{n\to\infty}X_n\cdot e_1/n>0)>0$. We will adapt the arguments in \cite{Ze} and prove that with positive probability, the number of visits to the $i$-th level $\mathcal{H}_i=\mathcal{H}_i(X_0):=\{x:x\cdot e_1=X_0\cdot e_1+ i\}$ grows slower than $Ci^2$. An important ingredient of the proof is a refinement of a combinatorial lemma of Zerner \cite[Lemma 3]{Ze} about deterministic paths. We say that a sequence $\{x_i\}_{i=0}^{k-1}\in (\mathbb{Z}^d)^{k}$, $2\le k\le\infty$, is a \textit{path} if $|x_i-x_{i-1}|=1$ for $i=1,\cdots, k-1$. For $i\ge 0$ and an infinite path $X_\cdot=\{X_n\}_{n=0}^\infty$ such that $\sup_n X_n\cdot e_1=\infty$, let \[T_i=\inf\{n\ge 0: X_n\in\mathcal{H}_i\}.\] For $0\le i<j$ and $k\ge 1$, let $T_{i,j}^1:=T_i$ and define recursively \[ T_{i,j}^{k+1}=\inf\{n\ge T_{i,j}^k: X_n\in\mathcal{H}_i \text{ and } n<T_j\}\in \mathbb{N}\cup \{\infty\}. \] That is, $T_{i,j}^k$ is the time of the $k$-th visit to $\mathcal{H}_i$ before hitting $\mathcal{H}_j$. Let \[ N_{i,j}=\sup\{k: T_{i,j}^k<\infty\} \] be the total number of visits to $\mathcal{H}_i$ before hitting $\mathcal{H}_j$. As in \cite{Ze}, for $i\ge 0, l\ge 1$, let \[ h_{i,l}=T_{i,i+l}^{N_{i,i+l}}-T_i \] denote the time spent between the first and the last visits to $\mathcal{H}_i$ before hitting $\mathcal{H}_{i+l}$. For $m,M, a\ge 0$ and $l\ge 1$, set \[ H_{m,l}=\sum_{i=0}^{l-1}N_{m+i,m+l}/(i+1)^2 \] and \[ E_{M,l}(a)=\frac{\#\{0\le m\le M: h_{m,l}\le a \text{ and } H_{m,l}\le a\}}{M+1}. \] Note that $E_{M,l}(a)$ decreases in $l$ and increases in $a$. The following lemma is a minor adaptation of \cite[Lemma 3]{Ze}. \begin{lemma}\label{LVl5} For any path $X_\cdot$ with $\varlimsup_{n\to\infty}X_n\cdot e_1/n>0$, \begin{equation}\label{LVe27} \sup_{a\ge 0}\inf_{l\ge 1}\varlimsup_{M\to\infty}E_{M,l}(a)>0. \end{equation} \end{lemma} {\it Proof:}~ Since $\varlimsup_{n\to\infty}n/T_n=\varlimsup_{n\to\infty}X_n\cdot e_1/n>0$, there exist an increasing sequence $(n_k)_{k=0}^\infty$ and $\delta<\infty$ such that \[ T_{n_k}<\delta n_k \text{ for all }k. \] Thus for any $m$ such that $n_k/2\le m\le n_k$, \begin{equation}\label{LV*17} T_m\le 2\delta m. \end{equation} Set $M_k=\lceil n_k/2\rceil$, where $\lceil x\rceil\in\mathbb{N}$ denotes the smallest integer which is not smaller than $x$. Then for all $k$ and $1<l<\lfloor n_k/2 \rfloor$, \begin{align}\label{LVe28} \sum_{m=0}^{M_k} H_{m,l} &=\sum_{i=0}^{l-1}\Big(\sum_{m=0}^{M_k}N_{m+i,m+l}\Big)/(i+1)^2\nonumber\\ &\le \sum_{i=0}^{l-1} T_{M_k+l}/(i+1)^2 \stackrel{(\ref{LV*17})}{\le} 4\delta (M_k+l). \end{align} By the same argument as in Page 193-194 of \cite{Ze}, we will show that there exist constants $c_1, c_2>0$ such that \begin{equation}\label{LV*18} \inf_{l\ge 1}\varlimsup_{k\to\infty} \frac{\#\{0\le m\le M_k: h_{m,l}\le c_1 \}}{M_k+1}>c_2. \end{equation} Indeed, if (\ref{LV*18}) fails, then for any $u>0$, \[ \varlimsup_{k\to\infty}\dfrac{\#\{0\le m\le M_k, h_{m,l}\le u\}}{M_k+1}\longrightarrow 0 \] as $l\to\infty$ (note that the right side is decreasing in $l$). Hence, one can find a sequence $(l_i)_{i\ge 0}$ with $l_{i+1}>l_i, l_0=0,$ such that for all $i\ge 0$, \begin{equation}\label{LV*19} \varlimsup_{k\to\infty}\dfrac{\#\{0\le m\le M_k, h_{m,l_{i+1}}\le 6\delta l_i\}}{M_k+1}<\frac{1}{3}. \end{equation} On the other hand, for $i\ge 0$ \begin{align} &\varlimsup_{k\to\infty}\dfrac{\#\{0\le m\le M_k, h_{m,l_i}\ge 6\delta l_i\}}{M_k+1}\nonumber\\ &\le \varlimsup_{k\to\infty}\frac{1}{(M_k+1)6\delta l_i}\sum_{m=0}^{M_k}(T_{m+l_i}-T_m)\nonumber\\ &\le \varlimsup_{k\to\infty}\frac{l_i T_{M_k+l_i}}{6\delta l_i(M_k+1)} \stackrel{(\ref{LV*17})}{\le}\frac{1}{3}. \label{LV*20} \end{align} By (\ref{LV*19}) and (\ref{LV*20}) , for any $i\ge 0$, \begin{equation}\label{LV*21} \varlimsup_{k\to\infty}\dfrac{\#\{0\le m\le M_k, h_{m,l_{i+1}}> h_{m,l_i}\}}{M_k+1} \ge \frac{1}{3}. \end{equation} Therefore, for any $j\ge 1$, noting that \[ \sum_{i=0}^{j-1}1_{h_{m,l_{i+1}}>h_{m,l_i}}\le N_{m,m+l_j}\le H_{m,l_j}, \] we have \begin{align*} \frac{j}{3}&\stackrel{(\ref{LV*21})}{\le} \varlimsup_{k\to\infty} \sum_{i=0}^{j-1}\dfrac{\#\{0\le m\le M_k, h_{m,l_{i+1}}> h_{m,l_i}\}}{M_k+1}\\ &\le\varlimsup_{k\to\infty} \frac{1}{M_k+1}\sum_{m=0}^{M_k}H_{m,l_j}\stackrel{(\ref{LVe28})}{\le} 4\delta, \end{align*} which is a contradiction if $j$ is large. This proves (\ref{LV*18}). It follows from (\ref{LV*18}) that, for any $l\ge 1$, there is a subsequence $(M'_k)$ of $(M_k)$ such that \[ \frac{\#\{0\le m\le M'_k: h_{m,l}\le c_1 \}}{M'_k+1}>c_2 \] for all $k$. Letting $c_3=9\delta/c_2$, we have that when $k$ is large enough, \[ \frac{1}{M'_k+1}\sum_{m=0}^{M'_k}1_{h_{m,l}\le c_1, H_{m,l}>c_3} \le \frac{1}{c_3(M'_k+1)}\sum_{m=0}^{M'_k}H_{m,l} \stackrel{(\ref{LVe28})}{\le} \frac{c_2}{2}. \] Hence for any $l>1$ and large $k$, \begin{align*} E_{M_k',l}(c_1\vee c_3) &\ge \frac{1}{M'_k+1}\sum_{m=0}^{M'_k}1_{h_{m,l}\le c_1,H_{m,l}\le c_3}\\ &= \frac{1}{M'_k+1}\sum_{m=0}^{M'_k}(1_{h_{m,l}\le c_1}-1_{h_{m,l}\le c_1,H_{m,l}> c_3}) \ge \frac{c_2}{2}. \end{align*} This shows the lemma, and what is more, with explicit constants.\qed\\ For $i\ge 0$, let $N_i=\mathop{\rm lim}_{j\to\infty} N_{i,j}$ denote the total number of visits to $\mathcal{H}_i$. With Lemma \ref{LVl5}, one can deduce that with positive probability, $N_i\le C(i+1)^2$ for all $i\ge 0$: \begin{theorem}\label{LVthm3} If $\mathbb{P}(\varlimsup_{n\to\infty}X_n\cdot e_1/n>0)>0$, then there exists a constant $c_5$ such that \[\mathbb{P}(R=\infty)>0,\] where $R$ is the stopping time defined by \begin{align*} R&=R_{e_1}(X_\cdot, c_5)\\ &:= \inf\{n\ge 0: \sum_{i=0}^n 1_{X_i\in\mathcal{H}_j}>c_5(j+1)^2 \text{ for some }j\ge 0\}\wedge D, \end{align*} and $D:=\inf\{n\ge 1: X_n\cdot e_1\le X_0\cdot e_1\}$. \end{theorem} \begin{figure}[h] \centering \includegraphics[width=0.7\textwidth]{intersection} \caption{On $\{R=\infty\}$, the path visits the $i$-th level no more than $c_5(i+1)^2$ times.} \end{figure} Note that for any $L>0$ and a path $(X_i)_{i=0}^\infty$ with $X_0=o$, \begin{align}\label{LVe32} \sum_{\substack{y: y\cdot e_1\le -L\\0\le i\le R}}e^{-\gamma d(y,X_i)} &\le \sum_{j=0}^\infty (\#\text{visits to $\mathcal{H}_j$ before time $R$})e^{-\gamma (j+L)}\nonumber\\ &\le C\sum_{j=0}^\infty c_5(j+1)^2e^{-\gamma(j+L)}\le Ce^{-\gamma L}. \end{align} Hence on the event $\{R=\infty\}$, by (\ref{LVe32}) and $(G)$, the trajectory $(X_i)_{i=0}^\infty$ is ``almost independent" with the environments $\{\omega_x:x\cdot e_1\le -L\}$ when $L$ is large. This fact will be used in our definition of the regeneration times in the Section \ref{seclln}. To prove Theorem \ref{LVthm3}, we need the following lemma. Recall that $r,\gamma$ are parameters of the environment measure $P$. Let $S$ be a countable set of finite paths. With abuse of notation, we also use $S$ as the synonym for the event \begin{equation}\label{LV2e*} \bigcup_{(x_i)_{i=0}^N\in S}\{X_i=x_i \text{ for }0\le i\le N\}. \end{equation} \begin{lemma}\label{LVc2} Let $a>0$ and $A\subset\Lambda\subset\mathbb{Z}^d$. Suppose $S\neq\emptyset$ is a countable set of finite paths $x_\cdot=(x_i)_{i=0}^N, N<\infty$ that satisfy $d(x_\cdot, \Lambda)\ge r$ and \[ \sum_{y\in A, 0\le i\le N}e^{-\gamma d(y,x_i)}\le a. \] Then, $P$-almost surely, \begin{equation}\label{LVe31} \exp(-Ca)\le\frac{E_P [P_\omega(S)|\omega_x: x\in\Lambda]}{E_P [P_\omega(S)|\omega_x: x\in\Lambda\setminus A]}\le\exp(Ca). \end{equation} \end{lemma} {\it Proof:}~ We shall first show that for any $(x_i)_{i=0}^N\in S$, $P$-almost surely, \begin{align}\label{LVe40}&E_P[P_\omega(X_i=x_i,0\le i\le N)|\omega_y:y\in\Lambda]\nonumber\\&\le \exp(Ca)E_P[P_\omega(X_i=x_i,0\le i\le N)|\omega_y:y\in\Lambda\setminus A].\end{align} Note that when $\Lambda^c$ is a finite subset of $\mathbb{Z}^d$, (\ref{LVe40}) is an easy consequence of $(G)$. For general $\Lambda$, we let \[ \Lambda_n=\Lambda\cup\{x:|x|\ge n\}. \] When $n$ is sufficiently big, $(G)$ implies that \begin{equation*} \frac{E_P[P_\omega(X_i=x_i,0\le i\le N)|\omega_y:y\in\Lambda_n]}{E_P[P_\omega(X_i=x_i,0\le i\le N)|\omega_y:y\in\Lambda_n\setminus A]}\le\exp(Ca). \end{equation*} Since $\Lambda_n\downarrow \Lambda$ as $n\to\infty$, (\ref{LVe40}) follows by taking $n\to\infty$ in the above inequality. Summing over all $(x_i)_{i=0}^N\in S$ on both sides of (\ref{LVe40}), we conclude that $P$-almost surely, \[ E_P[P_\omega(S)|\omega_y:y\in\Lambda] \le \exp(Ca)E_P[P_\omega(S)|\omega_y:y\in\Lambda\setminus A]. \] The upper bound of (\ref{LVe31}) is proved. The lower bound follows likewise.\qed Now we can prove the theorem. Our proof is a modification of the proof of Theorem 1 in \cite{Ze}:\\ \noindent\textit{Proof of Theorem \ref{LVthm3}:} It follows by Lemma \ref{LVl5} that there exists a constant $c_4>0$ such that \begin{equation}\label{LV*22} \mathbb{P}(\inf_{l\ge 1}\varlimsup_{M\to\infty}E_{M,l}(c_4)>0)>0. \end{equation} For $l>r$, $k\ge 0$ and $z\in\mathbb{Z}^d$ with $z\cdot e_1=r$, let $B_{m,l}(z,k,c)$ denote the event \[ \{N_{m+r,m+l}=k,X_{T_{m+r,m+l}^k}=X_{T_m}+z,H_{m+r,l-r}\le c\}. \] Note that on the event $\{h_{m,l}\le c_4\text{ and }H_{m,l}\le c_4\}$, we have \begin{align*} T_{m+r,m+l}^{N_{m+r,m+l}}-T_m &\le h_{m,l}+\sum_{i=0}^r N_{m+i,m+l}\\ &\le c_4+\sum_{i=0}^r (i+1)^2c_4\le (1+r)^3c_4, \shortintertext{and} H_{m+r,l-r} &\le \sum_{i=0}^{l-r-1}(r+1)^2N_{m+r+i,m+l}/(r+i+1)^2\\ &\le (r+1)^2c_4=:c_5. \end{align*} Hence $ \{h_{m,l}\le c_4\text{ and }H_{m,l}\le c_4\}\subset \bigcup_{|z|,k\le (r+1)^3c_4}B_{m,l}(z,k,c_5), $ and \[ \mathop{\rm lim}_{l\to\infty}\varlimsup_{M\to\infty}E_{M,l}(c_4) \le \sum_{|z|,k\le (r+1)^3c_4} \varlimsup_{l\to\infty}\varlimsup_{M\to\infty}\frac{1}{M+1} \sum_{m=0}^M 1_{B_{m,l}(z,k,c_5)}. \] Thus by (\ref{LV*22}), for some $k_0$ and $z_0$ with $z_0\cdot e_1=r$, \begin{equation}\label{LVe29} \mathbb{P}(\varlimsup_{l\to\infty}\varlimsup_{M\to\infty} \frac{1}{M+1}\sum_{m=0}^M 1_{B_{m,l}(z_0,k_0,c_5)}>0)>0. \end{equation} In what follows, we write $B_{m,l}(z_0,k_0,c_5)$ simply as $B_{m,l}$. For any $l>r$ and any fixed $i\le l-1$, let $m_j=m_j(l,i):=i+jl$, i.e. $(m_j)_{j\ge 0}$ is the class of residues of $i(\text{mod }l)$. Now take any $j\in \mathbb{N}$. Observe that for any event $E=\{1_{B_{m_{j-1},l}}=\cdot,\ldots,1_{B_{m_0,l}}=\cdot\}$ and $x\in \mathcal{H}_{m_j}$, \begin{align}\label{LV*1} \MoveEqLeft P_\omega(\{X_{T_{m_j}}=x\}\cap E\cap B_{m_j,l})\\ &\le P_\omega(\{X_{T_{m_j}}=x\}\cap E) P_\omega^{x+z_0}(D>T_{l-r},H_{0,l-r}\le c_5).\nonumber \end{align} Moreover, for any $x\in \mathcal{H}_{m_j}$, there exists a countable set $S$ of finite paths $(x_i)_{i=0}^N$ that satisfy $m_j+r\le x_i\cdot e_1\le m_j+l$ and $\#\{k\le N: x_k\in\mathcal{H}_i(x_0)\}\le c_5(i+1)^2$ for $0\le i\le N$, such that \begin{align*} &\{X_0=x+z_0, D>T_{l-r},H_{0,l-r}\le c_5\}\\ &=\cup_{(x_i)_{i=0}^N\in S}\{X_i=x_i \text{ for }0\le i\le N\}. \end{align*} Noting that (by the same argument as in (\ref{LVe32})) for any $(x_i)_{i=0}^N\in S$, \[ \sum_{\substack{y:y\cdot e_1\le m_j\\i\le N}}e^{-\gamma d(y,x_i)}\le Ce^{-\gamma r}, \] by Lemma \ref{LVc2} we have \begin{align*} &E_P[P_\omega^{x+z_0}(D>T_{l-r},H_{0,l-r}\le c_5)|\omega_y:y\cdot e_1\le m_j]\\ &\le \exp{(Ce^{-\gamma r})}\mathbb{P}(D>T_{l-r},H_{0,l-r}\le c_5). \end{align*} Thus for $j\ge 0$ and $l>r$, \begin{align*} &\mathbb{P}(E\cap B_{m_j,l})\\ &\stackrel{(\ref{LV*1})}{\le} \sum_{x\in\mathcal{H}_{m_j}} E_P \big[P_\omega(\{X_{T_{m_j}}=x\}\cap E)P_\omega^{x+z_0}(D>T_{l-r},H_{0,l-r}\le c_5)\big]\\ &\le \exp{(Ce^{-\gamma r})} \sum_{x\in\mathcal{H}_{m_j}} \mathbb{P}(\{X_{T_{m_j}}=x\}\cap E)\mathbb{P}(D>T_{l-r},H_{0,l-r}\le c_5)\\ &= C\mathbb{P}(E) \mathbb{P}(D>T_{l-r},H_{0,l-r}\le c_5). \end{align*} Hence, for any $j\ge 0$ and $l>r$, \begin{equation*} \mathbb{P}(1_{B_{m_j,l}}=1|1_{B_{m_{j-1},l}},\ldots,1_{B_{m_0,l}}) \le C\mathbb{P}(D>T_{l-r},H_{0,l-r}\le c_5), \end{equation*} which implies that $\mathbb{P}$-almost surely, \begin{equation}\label{LVe30} \varlimsup_{n\to\infty} \frac{1}{n}\sum_{j=0}^{n-1} 1_{B_{m_j,l}} \le C\mathbb{P}(D>T_{l-r},H_{0,l-r}\le c_5). \end{equation} Therefore, $\mathbb{P}$-almost surely, \begin{align*} \varlimsup_{l\to\infty}\varlimsup_{M\to\infty} \frac{1}{M+1}\sum_{m=0}^M 1_{B_{m,l}} &\le \varlimsup_{l\to\infty}\frac{1}{l}\sum_{i=0}^{l-1} \varlimsup_{M\to\infty}\frac{l}{M+1} \sum_{\substack{0\le m\le M\\m\text{ mod }l=i}} 1_{B_{m,l}}\\ &\stackrel{(\ref{LVe30})}{\le} \mathop{\rm lim}_{l\to\infty} C\mathbb{P}(D>T_{l-r},H_{0,l-r}\le c_5)\\ &=C\mathbb{P}(D=\infty, \sum_{i=0}^\infty N_i/(i+1)^2\le c_5). \end{align*} This and (\ref{LVe29}) yield $\mathbb{P}(D=\infty, \sum_{i=0}^\infty N_i/(i+1)^2\le c_5)>0$. The theorem follows.\qed \section{The conditional law of large numbers}\label{seclln} In this section we will prove the conditional law of large numbers \eqref{ICLLN}, using regeneration times and coupling. Given the dependence structure of the environment, we want to define regeneration times in such a way that what happens after a regeneration time has little dependence on the past. To this end, we will use the ``$\epsilon$-coins" trick introduced in \cite{CZ1} and the stopping time $R$ to define the regeneration times. Intuitively, at a regeneration time, the past and the future movements have nice properties. That is, the walker has walked straight for a while without paying attention to the environment, and his future movements have little dependence on his past movements. We define the $\epsilon$-coins $(\epsilon_{i,x})_{i\in\mathbb{N}, x\in \mathbb{Z}^d}=:\epsilon$ to be iid random variables with distribution $Q$ such that \[Q(\epsilon_{i,x}=1)=d\kappa \text{ and }Q(\epsilon_{i,x}=0)=1-d\kappa.\] For fixed $\omega$, $\epsilon$, $P_{\omega,\epsilon}^x$ is the law of the Markov chain $(X_n)$ such that $X_0=x$ and that for any $e\in\mathbb{Z}^d$ such that $|e|=1$, \[ P_{\omega,\epsilon}^x(X_{n+1}=z+e|X_n=z) =\frac{1_{\epsilon_{n,z}=1}}{2d}+\frac{1_{\epsilon_{n,z}=0}}{1-d\kappa}[\omega(z,z+e)-\frac{\kappa}{2}]. \] Note that the law of $X_\cdot$ under $\bar{P}_\omega^x=Q\otimes P_{\omega,\epsilon}^x$ coincides with its law under $P_\omega^x$. Sometimes we also refer to $P_{\omega,\epsilon}^x (\cdot)$ as a measure on the sets of paths, without indicating the specific random path. Denote by $\bar{\mathbb{P}}=P\otimes Q\otimes P_{\omega,\epsilon}^o$ the law of the triple $(\omega, \epsilon, X_\cdot)$. Now we define the regeneration times in the direction $e_1$. Let $L$ be a fixed number which is sufficiently large. Set $R_0=0$. Define inductively for $k\ge 0$: \begin{align*} &S_{k+1}=\inf\{n\ge R_k: X_{n-L}\cdot e_1>\max\{X_m\cdot e_1: m<n-L\},\\ &\qquad\qquad\qquad \epsilon_{n-i, X_{n-i}}=1, X_{n-i+1}-X_{n-i}=e_1 \text{ for all }1\le i\le L\},\\ &R_{k+1}=R\circ\theta_{S_{k+1}}+S_{k+1}, \end{align*} where $\theta_n$ denotes the time shift of the path, i.e., $\theta_n X=(X_{n+i})_{i=0}^\infty$. Let \[K=\inf\{k\ge 1: S_k<\infty,R_k=\infty\}\] and $\tau_1=\tau_1(e_1,\epsilon,X_\cdot):=S_K.$ For $k\ge 1$, the ($L$-)regeneration times are defined inductively by \[\tau_{k+1}=\tau_1\circ\theta_{\tau_k}+\tau_k .\] By similar argument as in \cite[Lemma 2.2]{CZ1}, we can show: \begin{lemma}\label{LVl7} If $\mathbb{P}(\mathop{\rm lim}_{n\to\infty}X_n\cdot e_1/n=0)<1$, then \begin{equation}\label{LVe19} \mathbb{P}(A_{e_1}\cup A_{-e_1})=1. \end{equation} Moreover, on $A_{e_1}$, $\tau_i$'s are $\bar{\mathbb P}$-almost surely finite. \end{lemma} {\it Proof:}~ If $\mathbb{P}(\mathop{\rm lim}_{n\to\infty}X_n\cdot e_1/n=0)<1$, \[ \mathbb{P}(\varlimsup_{n\to\infty}X_n\cdot e_1/n>0)>0\quad\text{ or }\quad \mathbb{P}(\varlimsup_{n\to\infty}X_n\cdot (-e_1)/n>0)>0. \] Without loss of generality, assume that \[\mathbb{P}(\varlimsup_{n\to\infty}X_n\cdot e_1/n>0)>0.\] It then follows from Theorem \ref{LVthm3} that $\mathbb{P}(R=\infty)>0$. We want to show that $R_k=\infty$ for all but finitely many $k$'s. For $k\ge 0$, \begin{align*} &\bar{\mathbb P}(R_{k+1}<\infty)\\ &= \bar{\mathbb P}(S_{k+1}<\infty, R\circ\theta_{S_{k+1}}<\infty)\\ &=\sum_{n,x}\bar{\mathbb P} (S_{k+1}=n, X_n=x, R\circ\theta_n <\infty)\\ &=\sum_{n,x}E_{P\otimes Q} \big[P_{\omega,\epsilon}(S_{k+1}=n,X_n=x) P_{\omega,\theta^n\epsilon}^x(R<\infty)\big], \end{align*} where $\theta^n\epsilon$ denotes the time shift of the coins $\epsilon$, i.e. $(\theta^n\epsilon)_{i,x}=\epsilon_{n+i,x}$. Note that $P_{\omega,\epsilon}(S_{k+1}=n,X_n=x)$ and $P_{\omega,\theta^n\epsilon}^x(R<\infty)$ are independent under the measure $Q$, since the former is a function of $\epsilon$'s before time $n$, and the latter involves $\epsilon$'s after time $n$. It then follows by induction that \begin{align*} &\bar{\mathbb P}(R_{k+1}<\infty)\\ &=\sum_{n,x}E_P \big[\bar{P}_\omega(S_{k+1}=n,X_n=x) \bar{P}_\omega^x(R<\infty)\big]\\ &=\sum_{n,x}E_P\big[\bar{P}_\omega(S_{k+1}=n,X_n=x) E_P[\bar{P}_\omega^x(R<\infty)|\omega_y: y\cdot e_1\le x\cdot e_1-L] \big]\\ &\stackrel{(\ref{LVe32}), \text{Lemma }\ref{LVc2}}{\le} \bar{\mathbb P}(R_k<\infty)\exp{(e^{-cL})}\bar{\mathbb P}(R<\infty)\\ &\le [\exp{(e^{-cL})}\bar{\mathbb P}(R<\infty)]^{k+1}, \end{align*} where we used in the second equality the fact that $\bar{P}_\omega(S_{k+1}=n,X_n=x)$ is $\sigma(\omega_y: y\cdot e_1\le x\cdot e_1-L)$-measurable. Hence, by taking $L$ sufficiently large and by the Borel-Cantelli Lemma, $\bar{\mathbb P}$-almost surely, $R_k=\infty$ except for finitely many values of $k$. Let $\mathcal{O}_{e_1}$ denote the event that the signs of $X_n\cdot e_1$ change infinitely many often. It is easily seen that (by the ellipticity of the environment) \begin{align*} &\mathbb{P}(\mathcal{O}_{e_1}\cup A_{e_1}\cup A_{-e_1})=1 \shortintertext{and} &\mathcal{O}_{e_1}\subset \{\sup_n X_n\cdot e_1=\infty\}. \end{align*} However, on $\{\sup_n X_n\cdot e_1=\infty\}$, given that $R_k$ is finite, $S_{k+1}$ is also finite. Hence $\tau_1$ is $\bar{\mathbb P}$-almost surely finite on $\{\sup_n X_n\cdot e_1=\infty\}$, and so are the regeneration times $\tau_2,\tau_3\ldots$. Therefore, \[ \mathbb{P}(\mathcal{O}_{e_1})=\bar{\mathbb P}(\mathcal{O}_{e_1}\cap\{\tau_1<\infty\}). \] Since $\mathcal{O}_{e_1}\cap\{\tau_1<\infty\}=\emptyset$, we get $\mathbb{P}(\mathcal{O}_{e_1})=0$. This gives (\ref{LVe19}). \qed\\ When $\mathbb{P}(R=\infty)>0$, we let \[ \hat{\mathbb P}(\cdot):=\bar{\mathbb P}(\cdot|R=\infty). \] The following proposition is a consequence of Lemma \ref{LVc2}. \begin{proposition} Assume $\mathbb{P}(R=\infty)>0$. Let $l>r$ and $\Lambda\subset\{x:x\cdot e_1<-r\}.$ Then for any $A\subset\Lambda\cap\{x: x\cdot e_1<-l\}$ and $k\in\mathbb{N}$, \begin{equation}\label{LVprop1} \exp(-Ce^{-\gamma l}) \le \dfrac{E_P\big[\bar{P}_\omega\big((X_i)_{i=0}^{\tau_k}\in\cdot, R=\infty\big)|\omega_y:y\in\Lambda\setminus A]} {E_P\big[\bar{P}_\omega\big((X_i)_{i=0}^{\tau_k}\in\cdot, R=\infty\big)|\omega_y:y\in\Lambda]} \le \exp(Ce^{-\gamma l}). \end{equation} Furthermore, for any $k\in\mathbb{N}$ and $n\ge 0$, $\hat{\mathbb P}$-almost surely, \begin{equation}\label{LVprop2} \exp(-e^{-cL}) \le \frac{\hat{\mathbb P}\big((X_{\tau_n+i}-X_{\tau_n})_{i=0}^{\tau_{n+k}-\tau_n}\in\cdot|X_{\tau_n}\big)} {\hat{\mathbb P}\big((X_i)_{i=0}^{\tau_k}\in\cdot\big)} \le \exp(e^{-cL}). \end{equation} \end{proposition} {\it Proof:}~ First, we shall prove (\ref{LVprop1}). By the definition of the regeneration times, for any finite path $x_\cdot=(x_i)_{i=0}^N, N<\infty$, there exists an event $G_{x_\cdot}\in\sigma(\epsilon_{i,X_i},X_i: i\le N)$ such that $G_{x_\cdot}\subset\{R>N\}$ and \[ \{(X_i)_{i=0}^{\tau_k}=(x_i)_{i=0}^N, R=\infty\} =G_{x_\cdot}\cap\{R\circ\theta_N=\infty\}. \] (For example, when $k=1$, we let \[ G_{x_\cdot}=\bigcup_{j=1}^\infty \{(X_i)_{i=0}^N=(x_i)_{i=0}^N, S_j=N, R>N\}. \] Then $\{(X_i)_{i=0}^{\tau_1}=(x_i)_{i=0}^N, R=\infty\} =G_{x_\cdot}\cap\{R\circ\theta_N=\infty\}.$) For $n\in\mathbb{N}$, we let \[ E_n:=G_{x_\cdot}\cap\{R\circ\theta_N\ge n\}. \] Note that $E_n\in\sigma(\epsilon_{i,X_i},X_i:i\le N+n)$ can be interpreted (in the sense of (\ref{LV2e*})) as a set of paths with lengths $\le N+n$. Also note that $E_n\subset\{R>N+n\}$. Then by Lemma \ref{LVc2} and (\ref{LVe32}), we have \[ \exp(-Ce^{-\gamma l}) \le \dfrac{E_P\big[\bar{P}_\omega\big(E_n)|\omega_y:y\in\Lambda\setminus A]} {E_P\big[\bar{P}_\omega\big(E_n\big)|\omega_y:y\in\Lambda]} \le \exp(Ce^{-\gamma l}). \] (\ref{LVprop1}) follows by letting $n\to\infty$. Next, we shall prove (\ref{LVprop2}). Let $x\in\mathbb{Z}^d$ be any point that satisfies \[ \bar{\mathbb P}(X_{\tau_n}=x)>0. \] By the definition of the regeneration times, for any $m\in\mathbb{N}$, there exists an event $G_m^x\in\sigma\{\epsilon_{i,X_i},X_i: i\le m\}$ such that $\bar{P}_\omega(G_m^x)$ is $\sigma(\omega_y:y\cdot e_1\le x\cdot e_1-L)$-measurable, and \[ \{\tau_n=m,X_m=x, R=\infty\}=G^x_m\cap\{R\circ\theta_m=\infty\}. \] Thus \begin{align}\label{LV2e2} &\bar{\mathbb P}\big((X_{\tau_n+i}-X_{\tau_n})_{i=0}^{\tau_{n+k}-\tau_n}\in\cdot,X_{\tau_n}=x, R=\infty\big)\nonumber\\ &=\sum_{m}\bar{\mathbb P}\big((X_{\tau_n+i}-X_{\tau_n})_{i=0}^{\tau_{n+k}-\tau_n}\in\cdot, \tau_n=m,X_m=x,R=\infty\big)\nonumber\\ &= \sum_{m}E_P\big[\bar{P}_\omega(G_m^x) \bar{P}_\omega^x((X_i-x)_{i=0}^{\tau_k}\in\cdot,R=\infty)\big]\nonumber\\ &\stackrel{(\ref{LVprop1})}{\le} \exp(Ce^{-\gamma L})\sum_{m}\bar{\mathbb P}(G_m^x) \bar{\mathbb P}\big((X_i)_{i=0}^{\tau_k}\in\cdot, R=\infty\big). \end{align} On the other hand, \begin{align}\label{LV2e11} \bar{\mathbb P}(X_{\tau_n}=x, R=\infty) &=\sum_{m} E_P[\bar{P}_{\omega}(G^x_m)\bar{P}_\omega^x(R=\infty)]\nonumber\\ &\stackrel{(\ref{LVprop1})}{\ge}\exp(-Ce^{-\gamma L}) \sum_{m} \bar{\mathbb P}(G_m^x)\bar{\mathbb P}(R=\infty). \end{align} By (\ref{LV2e2}) and (\ref{LV2e11}), we have (note that $L$ is sufficiently big) \[ \hat{\mathbb P}\big((X_{\tau_n+i}-X_{\tau_n})_{i=0}^{\tau_{n+k}-\tau_n}\in\cdot|X_{\tau_n}=x\big) \le \exp(e^{-cL})\hat{\mathbb P}\big((X_i)_{i=0}^{\tau_k}\in\cdot\big). \] The right side of (\ref{LVprop2}) is proved. The left side of (\ref{LVprop2}) follows likewise. \qed The next lemma describes the dependency of a regeneration on its remote past. It is a version of Lemma 2.2 in \cite{CZ2}. (The denominator is omitted in the last equality in \cite[page 101]{CZ2}, which is corrected here, see the equality in (\ref{LV*8}).) Set $\tau_0=0$. Denote the truncated path between $\tau_{n-1}$ and $\tau_n-L$ by \[ P_n=(P_n^i)_{0\le i\le \tau_{n}-\tau_{n-1}-L}:=(X_{i+\tau_{n-1}}-X_{\tau_{n-1}})_{0\le i\le \tau_n-\tau_{n-1}-L}. \] Set \begin{align*} W_n &=(\omega_{x+X_{\tau_{n-1}}})_{x\in P_n}=:\omega_{X_{\tau_{n-1}}+P_n},\\ F_n &=X_{\tau_n}-X_{\tau_{n-1}},\\ J_n &=(P_n,W_n,F_n,\tau_n-\tau_{n-1}). \end{align*} For $i\ge 0$, let $h_{i+1}(\cdot|j_i,\ldots,j_1):=\hat{\mathbb{P}}(J_{i+1}\in\cdot|J_{i},\ldots,J_1)|_{J_{i}=j_i,\ldots,J_1=j_1}$ denote the transition kernel of $(J_n)$. Note that when $i=0$, $h_{i+1}(\cdot|j_i,\ldots,j_1)=h_1(\cdot|\emptyset)=\hat{\mathbb P}(J_1\in\cdot)$. \begin{lemma}\label{LVl4} Assume $\mathbb{P}(R=\infty)>0$, $0\le k\le n$. Then $\hat{\mathbb P}$-almost surely, \begin{equation}\label{LVe26} \exp{(-e^{-c(k+1)L})} \le \frac{h_{n+1}(\cdot|J_n,\ldots,J_1)}{h_{k+1}(\cdot|J_n,\ldots,J_{n-k+1})} \le \exp{(e^{-c(k+1)L})}. \end{equation} \end{lemma} {\it Proof:}~ For $j_m=(p_m,w_m,f_m,t_m),m=1,\ldots n$, let \begin{align*} &\bar{x}_m:=f_1+\cdots+f_m,\\ &\bar{t}_m:=t_1+\cdots+t_m,\\ &B_{p_1,\ldots,p_m}:=\{R=\infty, P_i=p_i\text{ for all }i=1,\ldots,m\},\\ \text{ and }\quad & \omega_{p_1,\ldots,p_m}:=(\omega_{\bar{x}_{i-1}+p_i})_{i=1}^m. \end{align*} First, we will show that for any $1\le k\le n$, \begin{equation}\label{LV*8} h_{k+1}(\cdot|j_k,\ldots,j_1) =\frac{E_P\big[\bar{P}_\omega^{\bar{x}_k}(J_1\in \cdot,R=\infty)|\omega_{p_1,\ldots,p_k}\big]} {E_P\big[\bar{P}_\omega^{\bar{x}_k}(R=\infty)|\omega_{p_1,\ldots,p_k}\big]} \Big|_{\omega_{p_1,\ldots,p_k}=(w_i)_{i=1}^k}. \end{equation} By the definition of the regeneration times, there exists an event \[ G_{p_1,\ldots,p_k}\in\sigma(X_{i+1},\epsilon_{i,X_i}, 0\le i\le \bar{t}_k-1) \] such that \begin{equation}\label{LV*7} B_{p_1,\ldots,p_k}=G_{p_1,\ldots,p_k}\cap\{R\circ\theta_{\bar{t}_k}=\infty\}. \end{equation} On the one hand, for any $\sigma(J_k,\ldots, J_1)$-measurable function $g(J_k,\ldots,J_1)$, \begin{align}\label{LV*15} &E_{\bar{\mathbb P}}\big[h_{k+1}(\cdot|J_k,\ldots,J_1)g(J_k,\ldots, J_1)1_{B_{p_1,\ldots,p_k}}\big]\nonumber\\ &=E_{\bar{\mathbb P}}\big[g1_{B_{p_1,\ldots,p_k}}1_{J_{k+1}\in \cdot}\big]\nonumber\\ &= E_P [g1_{B_{p_1,\ldots,p_k}}\bar{P}_\omega(J_{k+1}\in\cdot,B_{p_1,\ldots,p_k})]\nonumber\\ &\stackrel{(\ref{LV*7})}{=} E_P \big[g1_{B_{p_1,\ldots,p_k}}\bar{P}_\omega(G_{p_1,\ldots,p_k})\bar{P}_\omega^{\bar{x}_k}(J_1\in \cdot,R=\infty)\big]. \end{align} On the other hand, we also have \begin{align}\label{LV*16} & E_{\bar{\mathbb P}}\big[h_{k+1}(\cdot|J_k,\ldots,J_1) g(J_k,\ldots, J_1)1_{B_{p_1,\ldots,p_k}}\big]\nonumber\\ &= E_P \big[ h_{k+1}(\cdot|J_k,\ldots,J_1)g1_{B_{p_1,\ldots,p_k}} \bar{P}_\omega(B_{p_1,\ldots,p_k})\big]\nonumber\\ &\stackrel{(\ref{LV*7})}{=} E_P \big[ h_{k+1}(\cdot|J_k,\ldots,J_1)g1_{B_{p_1,\ldots,p_k}} \bar{P}_\omega(G_{p_1,\ldots,p_k})\bar{P}_\omega^{\bar{x}_k}(R=\infty)\big]. \end{align} Comparing (\ref{LV*15}) and (\ref{LV*16}) and observing that on $B_{p_1,\ldots,p_k}$, $\bar{P}_\omega(G_{p_1,\ldots,p_k})$ and all functions of $J_1,\ldots,J_k$ are $\sigma(\omega_y: y\in\bar{x}_{i-1}+p_i, i\le k)$-measurable , we obtain that on $B_{p_1,\ldots,p_k}$, $P$-almost surely, \begin{equation*} h_{k+1}(\cdot|J_k,\ldots,J_1) =\frac{E_P\big[\bar{P}_\omega^{\bar{x}_k}(J_1\in \cdot,R=\infty)|\omega_{\bar{x}_{i-1}+p_i},i\le k\big]} {E_P\big[\bar{P}_\omega^{\bar{x}_k}(R=\infty)|\omega_{\bar{x}_{i-1}+p_i},i\le k\big]}. \end{equation*} Noting that \[ B_{p_1,\ldots,p_k}\cap\{\omega_{p_1,\ldots,p_k}=(w_i)_{i=1}^k\} = \{J_i=j_i,1\le i\le k\}, \] (\ref{LV*8}) is proved. Next, we will prove the lower bound in \eqref{LVe26}. When $n\ge k\ge 1$, by formula (\ref{LV*8}) and (\ref{LVprop1}), we have \begin{align}\label{LV3e2} &h_{n+1}(\cdot| j_n,\ldots,j_1)\nonumber\\ &= \frac{E_P[\bar{P}_\omega^{\bar{x}_n}(J_1\in\cdot,R=\infty)|\omega_{p_1,\ldots,p_n}]} {E_P\big[\bar{P}_\omega^{\bar{x}_n}(R=\infty)|\omega_{p_1,\ldots,p_n}\big]}\bigg|_{\omega_{p_1,\ldots,p_n}=(w_i)_{i=0}^n}\nonumber\\ &\le \frac{\exp(Ce^{-\gamma(k+1)L}) E_P[\bar{P}_\omega^{\bar{x}_n}(J_1\in\cdot,R=\infty)|\omega_{\bar{x}_{i-1}+p_i},n-k+1\le i\le n]} {\exp(-Ce^{-\gamma(k+1)L}) E_P[\bar{P}_\omega^{\bar{x}_n}(R=\infty)|\omega_{\bar{x}_{i-1}+p_i},n-k+1\le i\le n]}\nonumber\\ &\qquad \big|_{\omega_{p_1,\ldots,p_n}=(w_i)_{i=0}^n}\nonumber\\ &= \exp(2Ce^{-\gamma(k+1)L}) \frac{E_P[\bar{P}_\omega^{\bar{x}_n-\bar{x}_{n-k}}(J_1\in\cdot,R=\infty)|\omega_{p_{n-k+1},\ldots, p_n}]} {E_P[\bar{P}_\omega^{\bar{x}_n-\bar{x}_{n-k}}(R=\infty)|\omega_{p_{n-k+1},\ldots, p_n}]}\nonumber\\ &\qquad \big|_{\omega_{p_{n-k+1},\ldots,p_n}=(w_i)_{i=n-k+1}^n}\nonumber\\ &\stackrel{(\ref{LV*8})}{=} \exp(2Ce^{-\gamma(k+1)L})h_{k+1}(\cdot|j_{n},\ldots,j_{n-k+1}), \end{align} where we used the translation invariance of the measure $P$ in the last but one equality. When $k=0$ and $n\ge 1$, by formula \eqref{LV*8} and \eqref{LVprop1}, \begin{align}\label{LV3e1} h_{n+1}(\cdot| j_n,\ldots,j_1) &\le \frac{\exp(Ce^{-\gamma L})E_P[\bar{P}_\omega^{\bar{x}_n}(J_1\in\cdot,R=\infty)]} {\exp(-Ce^{-\gamma L})E_P[\bar{P}_\omega^{\bar{x}_n}(R=\infty)]}\nonumber\\ &=\exp(2Ce^{-\gamma L})\hat{\mathbb P}(J_1\in\cdot)\nonumber\\ &=\exp(2Ce^{-\gamma L})h_1(\cdot|\emptyset). \end{align} When $k=n=0$, \eqref{LVe26} is trivial. Hence combining \eqref{LV3e2} and \eqref{LV3e1}, the lower bound in (\ref{LVe26}) follows as we take $L$ sufficiently big. The upper bound follows likewise.\qed \begin{lemma}\label{LVl6} Suppose that a sequence of non-negative random variables $(X_n)$ satisfies \[ a\le \frac{\, \mathrm{d} P(X_{n+1}\in\cdot|X_1,\ldots, X_n)}{\, \mathrm{d} \mu}\le b \] for all $n\ge 1$, where $a\le 1\le b$ are constants and $\mu$ is a probability measure. Let $m_\mu\le \infty$ be the mean of $\mu$. Then almost surely, \begin{equation}\label{LVe21} a m_\mu\le\varliminf_{n\to\infty} \frac{1}{n}\sum_{i=1}^n X_i \le \varlimsup_{n\to\infty} \frac{1}{n}\sum_{i=1}^n X_i \le b m_\mu. \end{equation} \end{lemma} Before giving the proof, let us recall the ``splitting representation" of random variables: \begin{proposition}\cite[Page 94]{Tho}\label{LVprop} Let $\nu$ and $\mu$ be probability measures. Let $X$ be a random variable with law $\nu$. If for some $a\in(0,1)$, \[ \frac{\, \mathrm{d}\nu}{\, \mathrm{d}\mu}\ge a, \] then, enlarging the probability space if necessary, we can find independent random variables $\Delta, \pi, Z$ such that \begin{itemize} \item[i)] $\Delta$ is Bernoulli with parameter $1-a$, i.e., $P(\Delta=1)=1-a$, $P(\Delta=0)=a$; \item[ii)] $\pi$ is of law $\mu$, and $Z$ is of law $(\nu-a\mu)/(1-a)$; \item[iii)] $X=(1-\Delta)\pi+\Delta Z$. \end{itemize} \end{proposition} \noindent{\it Proof of Lemma \ref{LVl6}}:\\ By Proposition \ref{LVprop}, enlarging the probability space if necessary, there are random variables $\Delta_i,\pi_i,Z_i,i\ge 1$, such that for any $i\in\mathbb{N}$, \begin{itemize} \item $\Delta_i$ is Bernoulli with parameter $(1-a)$, and $\pi_i$ is of law $\mu$; \item $\Delta_i, \pi_i$ and $Z_i$ are mutually independent; \item $(\Delta_i,\pi_i)$ is independent of $\sigma(\Delta_k,\pi_k,Z_k: k<i)$; \item $X_i=(1-\Delta_i)\pi_i+\Delta_i Z_i$. \end{itemize} Note that since $X_i$'s are supported on $[0,\infty)$, $\pi_i\ge 0$ and $Z_i\ge 0$ for all $i\in\mathbb{N}$. Thus by the law of large numbers, almost surely, \[ \varliminf_{n\to\infty}\frac{1}{n}\sum_{i=1}^n X_i\ge \mathop{\rm lim}_{n\to\infty}\frac{1}{n}\sum_{i=1}^n (1-\Delta_i)\pi_i=a m_\mu. \] This proves the first inequality of (\ref{LVe21}). If $m_\mu=\infty$, the last inequality of (\ref{LVe21}) is trivial. Assume that $m_\mu<\infty$. Let $(\tilde{\Delta}_i)_{i\ge 1}$ be an iid Bernoulli sequence with parameter $1-b^{-1}$ such that every $\tilde{\Delta}_i$ is independent of all the $X_n$'s. By a similar splitting procedure, we can construct non-negative random variables $\tilde{\pi}_i,\tilde{Z}_i, i\ge 1$, such that $(\tilde{\pi}_i)_{i\ge 1}$ are iid with law $\mu$, and \[ \tilde{\pi}_i=(1-\tilde{\Delta}_i)X_i+\tilde{\Delta}_i\tilde{Z}_i. \] Let $Y_i=(1-b^{-1}-\tilde{\Delta}_i)X_i 1_{X_i\le i}$, we will first show that \begin{equation}\label{LVe22} \mathop{\rm lim}_{n\to\infty}\frac{1}{n}\sum_{i=1}^n Y_i=0. \end{equation} By Kronecker's Lemma, it suffices to show that \[ \sum_{i=1}^\infty \frac{Y_i}{i} \text{ converges.} \] Observe that $(\sum_{i=1}^n Y_i/i)_{n\in\mathbb{N}}$ is a martingale sequence. Moreover, for all $n\in\mathbb{N}$, \begin{align*} E\big(\sum_{i=1}^n \frac{Y_i}{i}\big)^2 = \sum_{i=1}^n EY_i^2/i^2 &\le \sum_{i=1}^\infty EX_i^2 1_{X_i\le i}/i^2\\ &\le b \sum_{i=1}^\infty E\tilde{\pi}_i^2 1_{\tilde{\pi}_i\le i}/i^2\\ &= b\int_0^\infty x^2 (\sum_{i\ge x}\frac{1}{i^2})\, \mathrm{d}\mu\\ &\le C\int_0^\infty x\, \mathrm{d}\mu=Cm_\mu<\infty. \end{align*} By the $L^2$-martingale convergence theorem, $\sum Y_i/i$ converges a.s. and in $L^2$. This proves (\ref{LVe22}). Since \[ \sum_i P(Y_i\neq (1-b^{-1}-\tilde{\Delta}_i)X_i) \le \sum_i P(X_i>i) \le b \sum_i P(\pi_1>i)\le b m_\mu<\infty, \] by the Borel-Cantelli lemma, it follows from (\ref{LVe22}) that \[ \mathop{\rm lim}_{n\to\infty}\frac{1}{n}\sum_{i=1}^n (1-b^{-1}-\tilde{\Delta}_i)X_i=0, \text{a.s.}. \] Hence almost surely, \begin{equation*} m_\mu=\mathop{\rm lim}_{n\to\infty}\frac{1}{n}\sum_{i=1}^n \tilde{\pi}_i \ge \varlimsup_{n\to\infty}\frac{1}{n}\sum_{i=1}^n (1-\tilde{\Delta}_i)X_i =\varlimsup_{n\to\infty}\frac{1}{n}\sum_{i=1}^n b^{-1} X_i. \end{equation*} The last inequality of (\ref{LVe21}) is proved.\qed \begin{theorem}\label{LVlln} There exist two deterministic numbers $v_{e_1},v_{-e_1}\ge 0$ such that $\mathbb{P}$-almost surely, \begin{equation}\label{LVe25} \mathop{\rm lim}_{n\to\infty}\frac{X_n\cdot e_1}{n}=v_{e_1} 1_{A_{e_1}}-v_{-e_1}1_{A_{-e_1}}. \end{equation} Moreover, if $v_{e_1}>0$, then $E_{\hat{\mathbb P}}\tau_1<\infty$ and $\mathbb{P}(A_{e_1}\cup A_{-e_1})=1$. \end{theorem} {\it Proof:}~ We only consider the nontrivial case that $\mathbb{P}(\mathop{\rm lim} X_n\cdot e_1/n=0)<1$, which by Lemma \ref{LVl7} implies $\mathbb{P}(A_{e_1}\cup A_{-e_1})=1$. Without loss of generality, assume $\mathbb{P}(\varlimsup_{n\to\infty}X_n\cdot e_1/n>0)>0$. We will show that on $A_{e_1}$, \[ \mathop{\rm lim}_{n\to\infty}X_n\cdot e_1/n=v_{e_1}>0, \text{ $\mathbb{P}$-a.s..} \] By (\ref{LVprop2}) and Lemma \ref{LVl6}, we obtain that $\mathbb{P}(\cdot|A_{e_1})$-almost surely, \begin{align} \exp{(-e^{-cL})}E_{\hat{\mathbb P}}X_{\tau_1}\cdot e_1 &\le \varliminf_{n\to\infty}\frac{X_{\tau_n}\cdot e_1}{n}\nonumber\\ &\le \varlimsup_{n\to\infty}\frac{X_{\tau_n}\cdot e_1}{n} \le \exp{(e^{-cL})}E_{\hat{\mathbb P}}X_{\tau_1}\cdot e_1,\label{LVe23}\\ \exp{(-e^{-cL})}E_{\hat{\mathbb P}}\tau_1 &\le \varliminf_{n\to\infty}\frac{\tau_n}{n} \le \varlimsup_{n\to\infty}\frac{\tau_n}{n} \le \exp{(e^{-cL})}E_{\hat{\mathbb P}}\tau_1. \label{LVe24} \end{align} Note that (\ref{LVe23}), (\ref{LVe24}) hold even if $E_{\hat{\mathbb P}}X_{\tau_1}\cdot e_1=\infty$ or $E_{\hat{\mathbb P}}\tau_1=\infty$. But it will be shown later that under our assumption, both of them are finite. We claim that \begin{equation}\label{LVe5} E_{\hat{\mathbb P}}X_{\tau_1}\cdot e_1<\infty. \end{equation} To see this, let $\Theta:=\{i: X_{\tau_k}\cdot e_1=i \text{ for some }k\in\mathbb{N}\}$. Since $\tau_i$'s are finite on $A_{e_1}$, there exist (recall that $\tau_0=0$) a sequence $(k_n)_{n\in\mathbb{N}}$ such that $X_{\tau_{k_n}}\cdot e_1\le n<X_{\tau_{k_n+1}}\cdot e_1$ for all $n\in\mathbb{N}$ and $\mathop{\rm lim}_{n\to\infty}k_n=\infty$. Hence for $n\ge 1$, \[ \frac{\sum_{i=1}^n 1_{i\in \Theta}}{n}\le \frac{k_n+1}{X_{\tau_{k_n}}\cdot e_1}, \quad\text{ $\hat{\mathbb P}$-a.s..} \] Then, $\hat{\mathbb P}$-a.s., \[ \varlimsup_{n\to\infty} \frac{\sum_{i=1}^n 1_{i\in \Theta}}{n} \le \varlimsup_{n\to\infty}\frac{n}{X_{\tau_n}\cdot e_1}. \] Let $B_k=\{\epsilon_{k,X_k}=0, X_{k+1}-X_k=e_1, \epsilon_{k+i,X_{k+i}}=1,X_{k+i+1}-X_{k+i}=e_1 \text{ for all }1\le i\le L\}$. Then \[ \bar{P}_\omega(B_k) \ge (d\kappa)^L(1-d\kappa)(\frac{\kappa}{2})(\frac{1}{2d})^L \stackrel{1\ge 2d\kappa}{>}(\frac{\kappa}{2})^{L+2}. \] Observe that by the definition of the regeneration times, for $n> L+1$, \begin{align*} &\{T_{n-L-1}=k,X_k= x-(L+1)e_1, R>k\}\cap B_k\cap\{R\circ\theta_{k+L+1}=\infty\}\\ &\subset\{R=\infty, n\in\Theta, T_n=k+L+1,X_{T_n}=x\}. \end{align*} Hence for $n> L+1$, \begin{align*} & \hat{\mathbb P}(n\in \Theta)\\ &\ge \sum_{k\in\mathbb{N},x\in\mathcal{H}_n} \hat{\mathbb P}(B_k\cap\{T_{n-L-1}=k,X_k= x-(L+1)e_1,R\circ\theta_{k+L+1}=\infty\})\\ &\ge \sum_{k\in\mathbb{N},x\in\mathcal{H}_n} E_P \big[P_\omega\big(T_{n-L-1}=k,X_k= x-(L+1)e_1,R>k\big)(\frac{\kappa}{2})^{L+2}\\ &\qquad\qquad\qquad\qquad\qquad\qquad\qquad\quad\times P_\omega^x(R=\infty)\big]/\mathbb{P}(R=\infty). \end{align*} Since by (\ref{LVprop1}) and the translation invariance of $P$, \[ E_P\big[P_\omega^x(R=\infty)|\omega_y:y\cdot e_1\le x\cdot e_1-L-1\big] \ge \exp(-e^{-cL})\mathbb{P}(R=\infty), \] we have for $n>L+1$, \begin{align}\label{LV2e10} & \hat{\mathbb P}(n\in \Theta)\nonumber\\ &\ge (\frac{\kappa}{2})^{L+2}\exp(-e^{-cL}) \sum_{k\in\mathbb{N},x\in\mathcal{H}_n}\mathbb{P}(T_{n-L-1}=k,X_k= x-(L+1)e_1,R>k)\nonumber\\ &\ge (\frac{\kappa}{2})^{L+2} e^{-1}\mathbb{P}(R=\infty). \end{align} Hence \begin{align*} \frac{C}{E_{\hat{\mathbb P}}X_{\tau_1}\cdot e_1} \stackrel{(\ref{LVe23})}{\ge} E_{\hat{\mathbb P}}\varlimsup_{n\to\infty}\frac{n}{X_{\tau_n}\cdot e_1} &\ge E_{\hat{\mathbb P}}\varlimsup_{n\to\infty} \frac{\sum_{i=1}^n 1_{i\in \Theta}}{n}\\ &\ge \varlimsup_{n\to\infty}E_{\hat{\mathbb P}}\frac{\sum_{i=1}^n 1_{i\in \Theta}}{n}\\ &\stackrel{(\ref{LV2e10})}{\ge} (\frac{\kappa}{2})^{L+2} e^{-1}\mathbb{P}(R=\infty)>0. \end{align*} This gives (\ref{LVe5}). Now we can prove the theorem. By (\ref{LVe23}) and (\ref{LVe24}), \begin{align}\label{LV*9} \exp{(-2e^{-cL})}\frac{E_{\hat{\mathbb P}}X_{\tau_1}\cdot e_1}{E_{\hat{\mathbb P}}\tau_1} &\le \varliminf_{n\to\infty}\frac{X_{\tau_n}\cdot e_1}{\tau_{n+1}}\nonumber\\ &\le \varlimsup_{n\to\infty}\frac{X_{\tau_{n+1}}\cdot e_1}{\tau_n} \le \exp{(2e^{-cL})}\frac{E_{\hat{\mathbb P}}X_{\tau_1}\cdot e_1}{E_{\hat{\mathbb P}}\tau_1}, \end{align} $\mathbb{P}(\cdot|A_{e_1})$-almost surely. Further, by the fact that $|X_i|\le i$ and the obvious inequalities \begin{equation*} \varliminf_{n\to\infty}\frac{X_{\tau_n}\cdot e_1}{\tau_{n+1}} \le \varliminf_{n\to\infty}\frac{X_n\cdot e_1}{n} \le \varlimsup_{n\to\infty}\frac{X_n\cdot e_1}{n} \le \varlimsup_{n\to\infty}\frac{X_{\tau_{n+1}}\cdot e_1}{\tau_n}, \end{equation*} we have that \begin{equation*} \varlimsup_{n\to\infty} \Bigl\lvert \frac{X_n\cdot e_1}{n}- \frac{E_{\hat{\mathbb P}}X_{\tau_1}\cdot e_1}{E_{\hat{\mathbb P}}\tau_1}\Bigr\rvert \le \exp{(2e^{-cL})}-1, \text{ $\mathbb{P}(\cdot|A_{e_1})$-a.s.} \end{equation*} Therefore, $\mathbb{P}(\cdot|A_{e_1})$-almost surely, \[ \mathop{\rm lim}_{n\to\infty} \frac{X_n\cdot e_1}{n} = \mathop{\rm lim}_{L\to\infty} \frac{E_{\hat{\mathbb P}}X_{\tau_1^{(L)}}\cdot e_1}{E_{\hat{\mathbb P}}\tau_1^{(L)}}:=v_{e_1}, \] where $\tau_1$ is written as $\tau_1^{(L)}$ to indicate that it is an $L$-regeneration time. Moreover, our assumption $\mathbb{P}(\varlimsup_{n\to\infty}X_n\cdot e_1/n>0)>0$ implies that $v_{e_1}>0$ and (by (\ref{LV*9})) \[ E_{\hat{\mathbb P}}\tau_1<\infty. \] Our proof is complete.\qed\\ If $v_{e_1}>0$, then it follows by (\ref{LVe24}) that \begin{equation}\label{LVetau} E_{\hat {\mathbb P}}\tau_n\le CnE_{\hat{\mathbb P}}\tau_1<\infty. \end{equation} Observe that although Theorem \ref{LVlln} is stated for $e_1$, the previous arguments, if properly modified, still work if one replaces $e_1$ with any $z\in\mathbb{R}^d\setminus\{o\}$. So Theorem \ref{LVlln} is true for the general case. That is, for any $z\neq o$, there exist two deterministic constants $v_z, v_{-z}\ge 0$ such that \[ \mathop{\rm lim}_{n\to\infty}\frac{X_n\cdot z}{n}=v_z1_{A_z}-v_{-z}1_{A_{-z}} \] and that $\mathbb{P}(A_z\cup A_{-z})=1$ if $v_z>0$. Then, by the same argument as in \cite[page 1112]{Go}, one concludes that the limiting velocity $\mathop{\rm lim}_{n\to\infty}X_n/n$ can take at most two antipodal values. This proves the first part of Theorem \ref{LVthm2}. \section{Heat kernel estimates}\label{sechke} The following heat kernel estimates are crucial for the proof of the uniqueness of the non-zero velocity in the next section. Although in the mixing case we don't have iid regeneration slabs, we know that (by Lemma \ref{LVl4}) a regeneration slab has little dependence on its remote past. This allows us to use coupling techniques to get the same heat kernel estimates as in \cite{Be}: \begin{theorem}\label{hke} Assume $v_{e_1}>0$. For $x\in\mathbb{Z}^d$ and $n\in \mathbb{N}$, we let \[Q(n,x):=\hat{\mathbb P}(x \text{ is visited in }[\tau_{n-1},\tau_n)).\] Then for any $x\in\mathbb{Z}^d$ and $n\in \mathbb{N}$, \begin{align} &\hat{\mathbb P}(X_{\tau_n}=x)\le Cn^{-d/2},\label{LVehke}\\ & \sum_{x\in\mathbb{Z}^d}Q(n,x)^2\le C(E_{\hat{\mathbb P}}\tau_1)^2 n^{-d/2}.\label{LVehke2} \end{align} \end{theorem} By Lemma \ref{LVl4}, we have for $n\ge 2$ and $1\le k\le n-1$, $\hat{\mathbb P}$-almost surely, \begin{align}\label{LVnew1} \frac{h_{k+1}(\cdot|J_{n-1},\ldots,J_{n-k})}{h_k(\cdot|J_{n-1},\ldots,J_{n-k+1})} &= \frac{h_{k+1}(\cdot|J_{n-1},\ldots,J_{n-k})}{h_n(\cdot|J_{n-1},\ldots,J_1)} \frac{h_n(\cdot|J_{n-1},\ldots,J_1)}{h_k(\cdot|J_{n-1},\ldots,J_{n-k+1})}\nonumber\\ &\ge\exp(-e^{-c(k+1)L}-e^{-ckL})\nonumber\\ &\ge 1-e^{-ckL} \end{align} for large $L$. Hence for $n\ge 2$ and $1\le k\le n-1$, we can define a (random) probability measure $\zeta_{n,k}^{J_{n-1},\ldots,J_{n-k}}$ that satisfies \begin{align}\label{LVnew2} \MoveEqLeft h_{k+1}(\cdot|J_{n-1},\ldots,J_{n-k})\\ &=e^{-ckL}\zeta_{n,k}^{J_{n-1},\ldots,J_{n-k}}(\cdot)+(1-e^{-ckL})h_k(\cdot|J_{n-1},\ldots, J_{n-k+1}).\nonumber \end{align} To prove Theorem \ref{hke}, we will first construct a sequence of random variables $(\tilde J_i, i\in\mathbb{N})$ such that for any $n\in\mathbb{N}$, \begin{equation}\label{LVnew0} (\tilde J_1,\ldots,\tilde J_n)\sim \hat{\mathbb P}(J_1\in\cdot,\ldots, J_n\in\cdot), \end{equation} where ``$X\sim\mu$" means ``$X$ is of law $\mu$". \subsection{Construction of the $\tilde J_i$'s} Our construction consists of three steps: \noindent{\it Step 1.} We let $\tilde J_1, \tilde J_{2,1}, \tilde \Delta_{2,1}$ be independent random variables such that \[ \tilde J_1\sim h_1(\cdot|\emptyset),\quad \tilde J_{2,1}\sim h_1(\cdot|\emptyset) \] and $\tilde \Delta_{2,1}$ is Bernoulli with parameter $e^{-cL}$. Let $\tilde Z_{2,1}$ be independent of $\sigma(\tilde J_{2,1}, \tilde \Delta_{2,1})$ such that \[ P(\tilde Z_{2,1}\in\cdot|\tilde J_1)= \zeta_{2,1}^{\tilde J_1}(\cdot). \] Setting $\tilde J_{2}:=(1-\tilde \Delta_{2,1})\tilde J_{2,1}+\tilde\Delta_{2,1}\tilde Z_{2,1}$, by \eqref{LVnew2} we have \[ (\tilde J_1, \tilde J_2)\sim \hat{\mathbb P}(J_1\in\cdot,J_2\in\cdot). \] \noindent{\it Step 2.} For $n\ge 3$, assume that we have constructed $\tilde J_1$ and $(\tilde J_{i,1}, \tilde\Delta_{i,j}, \tilde Z_{i,j}, 1\le j<i\le n-1)$ such that \[ (\tilde J_1,\ldots, \tilde J_{n-1}) \sim \hat{\mathbb P}(J_1\in\cdot,\ldots,J_{n-1}\in\cdot), \] where for $2\le j\le i<n$, \[ \tilde J_{i,j}:=(1-\tilde\Delta_{i,j-1})\tilde J_{i,j-1}+\tilde\Delta_{i,j-1}\tilde Z_{i,j-1} \] and \[ \tilde J_i:= \tilde J_{i,i}. \] Then we define $\tilde J_{n,1}$ and $(\tilde\Delta_{n,k}, \tilde Z_{n,k}, 1\le k\le n-1)$ to be random variables such that conditioning on the values of $\tilde J_1$ and $(\tilde J_{i,1}, \tilde\Delta_{i,j}, \tilde Z_{i,j}, 1\le j<i\le n-1)$, \begin{itemize} \item $(\tilde J_{n,1}, \tilde\Delta_{n,k}, \tilde Z_{n,k}, 1\le k\le n-1)$ are conditionally independent; \item The conditional distribution of $\tilde J_{n,1}$ is $h_1(\cdot|\emptyset)$; \item For $1\le k\le n-1$, the conditional distributions of $\tilde Z_{n,k}$ and $\tilde\Delta_{n,k}$ are $\zeta_{n,k}^{\tilde J_{n-1},\ldots, \tilde J_{n-k}}(\cdot)$ and Bernoulli with parameter $e^{-ckL}$, respectively. \end{itemize} \noindent{\it Step 3.} For $2\le k\le n$, set \begin{align*} &\tilde J_{n,k}:=(1-\tilde\Delta_{n,k-1})\tilde J_{n,k-1}+\tilde\Delta_{n,k-1}\tilde Z_{n,k-1}\\ \text{and }&\tilde J_n:=\tilde J_{n,n}. \end{align*} Then (by \eqref{LVnew2}) almost surely, \begin{equation}\label{LV3e3} P(\tilde J_{n,k}\in\cdot|\tilde J_{n-1},\ldots,\tilde J_1)=h_k(\cdot|\tilde J_{n-1},\ldots,\tilde J_{n-k+1}). \end{equation} It follows immediately that \begin{equation*} (\tilde J_1,\ldots,\tilde J_n)\sim \hat{\mathbb P}(J_1\in\cdot,\ldots, J_n\in\cdot). \end{equation*} Therefore, by induction, we have constructed $(\tilde J_i, i\in\mathbb{N})$ such that \eqref{LVnew0} holds for all $n\in\mathbb{N}$. In what follows, with abuse of notation, we will identify $\tilde J_i$ with $J_i$ and simply write $\tilde J_{i,j}, \tilde \Delta_{i,j}, \tilde Z_{i,j}$ as $J_{i,j}, \Delta_{i,j}$ and $Z_{i,j}$, $1\le j<i$. We still use $\hat{\mathbb P}$ to denote the law of the random variables in the enlarged probability space. \begin{remark} To summarize, we have introduced random variables $J_{i,j}, \Delta_{i,j}, Z_{i,j}$, $1\le j<i$ such that for any $n\ge 2$, \begin{align*} &J_{n,2}=(1-\Delta_{n,1})J_{n,1}+\Delta_{n,1}Z_{n,1},\\ &\ldots,\\ &J_{n,n-1}=(1-\Delta_{n,n-2})J_{n,n-2}+\Delta_{n,n-2}Z_{n,n-2},\\ &J_n=J_{n,n}=(1-\Delta_{n,n-1})J_{n,n-1}+\Delta_{n,n-1}Z_{n,n-1}. \end{align*} Intuitively, we flip a sequence of ``coins" $\Delta_{n,n-1},\ldots,\Delta_{n,1}$ to determine whether $J_1,\ldots,J_{n-1}$ are in the ``memory" of $J_n$. For instance, if \[ \Delta_{n,n-1}=\cdots=\Delta_{n,n-i}=0, \] then $J_n=J_{n,n-i}$ doesn't ``remember" $J_1,\ldots, J_i$ (in the sense that \[ \hat{\mathbb P}(J_{n,n-i}\in\cdot|J_{n-1},\ldots, J_1) = h_{n-i}(\cdot|J_{n-1},\ldots, J_{i+1}). \] See \eqref{LV3e3}). \end{remark} \subsection {Proof of Theorem \ref{hke}} For $1<i\le n$, let $I_n(i)$ be the event that $\Delta_{i,i-1}=\ldots=\Delta_{i,1}=0$ and $\Delta_{m,m-1}=\ldots=\Delta_{m,m-i}=0$ for all $i<m\le n$. Note that on $I_n(i)$, \begin{equation}\label{LVnew3} J_i=J_{i,1}\text{ and }J_m=J_{m,m-i} \text{ for all }i<m\le n. \end{equation} Setting \[M_n:=\{1\le i\le n: I_n(i)\neq\emptyset\},\] we have \begin{lemma}\label{LVliid} For $n\ge 2$, let $H$ be a nonempty subset of $\{2,\ldots, n\}$, and set \[M_n:=\{1< i< n: I_n(i)\neq\emptyset\}.\] Conditioning on the event $\{M_n=H\}$, the sequence $(J_i)_{i\in H}$ is iid and independent of $(J_i)_{i\in \{1,\ldots,n\}\setminus H}$. \end{lemma} \noindent{\it Proof of Lemma \ref{LVliid}:} From our construction it follows that for any $i>1$, $J_{i,1}$ is independent of \[ \sigma(\Delta_{k,j}, 1\le j<k) \vee \sigma(J_l, 1\le l<i) \vee \sigma(J_{m,m-i},m>i). \] Hence by \eqref{LVnew3}, for any $i\in H$ and any appropriate measurable sets $(V_j)_{1\le j\le n}$, \begin{align*} &\hat{\mathbb P}(J_j\in V_j, 1\le j\le n|M_n=H)\\ &=\hat{\mathbb P}(J_{i,1}\in V_i) \hat{\mathbb P}(J_j\in V_j, 1\le j\le n, j\neq i|M_n=H). \end{align*} By induction, we get \begin{align*} &\hat{\mathbb P}(J_j\in V_j, 1\le j\le n|M_n=H)\\ &=\prod_{i\in H}\hat{\mathbb P}(J_{i,1}\in V_i) \hat{\mathbb P}(J_j\in V_j, 1\le j\le n, j\notin H|M_n=H). \end{align*} The lemma is proved.\qed \noindent{\it Proof of Theorem \ref{hke}:} By Lemma \ref{LVliid}, for $i\in H$ and all $j\in\{1,\ldots, d\}$, \[ \hat{\mathbb P}\big(X_{\tau_i}-X_{\tau_{i-1}}=(L+1)e_1\pm e_j|M_n=H\big) = \hat{\mathbb P}(X_{\tau_1}=(L+1)e_1\pm e_j)>0, \] where the last inequality is due to ellipticity. Hence arguing as in \cite[pages 736, 737]{Be}, using Lemma~\ref{LVliid} and the heat kernel estimate for bounded iid random walks in $\mathbb{Z}^d$, we get that for any $x\in\mathbb{Z}^d$, \[ \hat{\mathbb P}(\sum_{i\in H}X_{\tau_i}-X_{\tau_{i-1}}=x|M_n=H) \le C|H|^{-d/2}, \] where $|H|$ is the cardinality of $H$. Hence, for any subset $H\subset\{2,\ldots,n\}$ such that $|H|\ge n/2$, \begin{align}\label{LVe13} & \hat{\mathbb P}(X_{\tau_n}=x|M_n=H)\nonumber\\ &=\sum_y \hat{\mathbb P}(\sum_{i\in H}X_{\tau_i}-X_{\tau_{i-1}}=x-y, \sum_{i\in \{1,\ldots,n\}\setminus H}X_{\tau_i}-X_{\tau_{i-1}}=y|M_n=H)\nonumber\\ &=\sum_y \bigg[\hat{\mathbb P}\big(\sum_{i\in H}X_{\tau_i}-X_{\tau_{i-1}}=x-y|M_n=H\big)\nonumber\\ &\qquad\qquad\qquad\qquad\qquad\times \hat{\mathbb P}(\sum_{i\in \{1,\ldots,n\}\setminus H}X_{\tau_i}-X_{\tau_{i-1}}=y|M_n=H)\bigg]\nonumber\\ &\le C n^{-d/2}, \end{align} where we used Lemma~\ref{LVliid} in the second equality. On the other hand, \begin{align*} |M_n| &\ge n-\sum_{i=2}^n \bigg( 1_{\Delta_{i,i-1}+\cdots+\Delta_{i,1}>0}+\sum_{m=i+1}^n 1_{\Delta_{m,m-1}+\cdots+\Delta_{m,m-i}>0} \bigg)\\ &=n-\sum_ {i=2} ^n 1_{\Delta_{i,i-1}+\cdots+\Delta_{i,1}>0}-\sum_{m=2}^n\sum_{i=2}^{m-1}1_{\Delta_{ m,m-1}+\cdots+\Delta_{m,m-i}>0}\\ &\ge n-2\sum_{m=2}^n K_m, \end{align*} where $K_m:=\sup\{1\le j<m:\Delta_{m,j}=1\}$. Here we follow the convention that $\sup\emptyset=0$. Since $K_m$'s are independent, and for $m\ge 2$, \begin{align*} E e^{K_m} &= \sum_{j=0}^{m-1} e^j \hat{\mathbf P}(K_m=j)\\ &\le \sum_{j=1}^{m-1} e^j \hat{\mathbf P}(\Delta_{m,j}=1)+1\\ &\le \sum_{j=1}^\infty e^j e^{-cjL}+1\to 1 \text{ as $L\to\infty$}, \end{align*} we can take $L$ to be large enough such that $E e^{K_m}\le e^{1/8}$ for all $m\ge 2$ and so \begin{align}\label{LVe14} \hat{\mathbb P}(|M_n|<n/2) &\le \hat{\mathbb P}(K_2+\cdots+K_n>n/4)\nonumber\\ &\le e^{-n/4}E e^{K_2+\cdots +K_n} \le e^{-n/8}. \end{align} By (\ref{LVe13}) and (\ref{LVe14}), inequality (\ref{LVehke}) follows immediately. Furthermore, since \begin{align*} & Q(n,x)\\ &= \sum_y \hat{\mathbb P}(X_{\tau_{n-1}}=y) \hat{\mathbb P}(x \text{ is visited in }[\tau_{n-1},\tau_n)|X_{\tau_{n-1}}=y)\\ &\stackrel{\text{Lemma }\ref{LVl4}}{\le} C\sum_y \hat{\mathbb P}(X_{\tau_{n-1}}=y)\hat{\mathbb P}((x-y)\text{ is visited during }[0,\tau_1)), \end{align*} by H\"{o}lder's inequality we have \begin{align*} & Q(n,x)^2\\ &\le C \big[\sum_y\hat{\mathbb P}\big((x-y)\text{ is visited during }[0,\tau_1)\big)\big]\\ &\qquad\qquad\times\big[\sum_y\hat{\mathbb P}(X_{\tau_{n-1}}=y)^2 \hat{\mathbb P}\big((x-y)\text{ is visited during }[0,\tau_1)\big)\big]\\ &\le CE_{\hat{\mathbb P}}\tau_1 \sum_y \hat{\mathbb P}(X_{\tau_{n-1}}=y)^2 \hat{\mathbb P}\big((x-y)\text{ is visited during }[0,\tau_1)\big). \end{align*} Hence \begin{align*} & \sum_x Q(n,x)^2\\ &\le CE_{\hat{\mathbb P}}\tau_1 \sum_y \big[\hat{\mathbb P}(X_{\tau_{n-1}}=y)^2 \sum_x\hat{\mathbb P}\big((x-y)\text{ is visited during }[0,\tau_1)\big)\big]\\ &\le C(E_{\hat{\mathbb P}}\tau_1)^2 \sum_y \hat{\mathbb P}(X_{\tau_{n-1}}=y)^2\\ &\stackrel{(\ref{LVehke})}{\le} C(E_{\hat{\mathbb P}}\tau_1)^2 n^{-d/2}\sum_y \hat{\mathbb P}(X_{\tau_{n-1}}=y) =C(E_{\hat{\mathbb P}}\tau_1)^2 n^{-d/2}. \end{align*} Theorem \ref{hke} is proved.\qed \section{The uniqueness of the non-zero velocity}\label{secunique} In this section we will show that in high dimensions ($d\ge 5$), there exists at most one non-zero velocity. The idea is the following. Consider two random walk paths: one starts at the origin, the other starts near the $n$-th regeneration position of the first path. By Levy's martingale convergence theorem, the second path is ``more and more transient" as $n$ grows (Lemma \ref{LVl1}). On the other hand, by heat kernel estimates, when $d\ge 5$, two ballistic walks in opposite directions will grow further and further apart from each other (see Lemma \ref{LVl3}), thus they are almost independent. This contradicts the previous fact that starting at the $n$-th regeneration point of the first path will prevent the second path from being transient in the opposite direction.\\ Set $\delta=\delta(d):=\frac{d-4}{8(d-1)}$. (The reason for choosing this constant will become clear in (\ref{LV2e6}).). For any finite path $y_\cdot=(y_i)_{i=0}^M, M<\infty$, define $A(y_\cdot, z)$ to be the set of paths $(x_i)_{i=0}^N, N\le\infty$ that satisfy \begin{itemize} \item[1)] $x_0=y_0+z$; \item[2)] $d(x_i,y_j)>(i\vee j)^\delta$ if $i\vee j>|z|/3$. \end{itemize} The motivation for the definition of $A(y_\cdot, z)$ is as follows. Note that for two paths $x_\cdot=(x_i)_{i=0}^N$ and $y_\cdot=(y_i)_{i=0}^M$ with $x_0=y_0+z$, if $i\vee j\le |z|/3$, then \[ d(x_i,y_j) \ge d(x_0,y_0)-d(x_0,x_i)-d(y_0,y_j) \ge |z|-i-j \ge |z|/3. \] Hence, for $(x_i)_{i=0}^N\in A(y_\cdot, z)$, \begin{align}\label{LVe10} \sum_{i\le N,j\le M}e^{-\gamma d(x_i,y_j)} &\le \sum_{0\le i,j\le |z|/3}e^{-\gamma |z|/3}+\sum_{i\vee j>|z|/3}e^{-\gamma(i^\delta+j^\delta)/2} \nonumber\\ &\le (\frac{|z|}{3})^2 e^{-\gamma |z|/3}+(\sum_{i=0}^\infty e^{-\gamma i^\delta/2})^2<C. \end{align} This gives us (by (G)) an estimate of the interdependence between $\sigma\big(\omega_x: x\in (x_i)_{i=0}^N\big)$ and $\sigma\big(\omega_x: x\in (y_i)_{i=0}^M\big)$. In what follows, we use \[ \tau'_\cdot=\tau_\cdot(-e_1,\epsilon,X_\cdot) \] to denote the regeneration times in the $-e_1$ direction. Assume that there are two opposite nonzero limiting velocities in directions $e_1$ and $-e_1$, i.e., \[ v_{e_1} \cdot v_{-e_1}>0. \] We let $\check{\mathbb P}(\cdot):=\mathbb{P}(\cdot|R_{-e_1}=\infty)$. \begin{figure}[h] \centering \includegraphics[width=0.6\textwidth]{paths2} \caption{$X_\cdot\in A(Y_\cdot^n, z)$. When $i\vee j>|z|/3$, the distance between $Y_j^n$ of the ``backward path" and $X_i$ is at least $(i\vee j)^\delta$.} \label{LVfig:1} \end{figure} \begin{lemma}\label{LVl3} Assume that there are two nonzero limiting velocities in direction $e_1$. We sample $(\epsilon,\tilde{X}_\cdot)$ according to $\hat{\mathbb P}$ and let $\tilde{\tau}_\cdot=\tau_\cdot (e_1,\epsilon,\tilde{X}_\cdot)$ denote its regeneration times. For $n\ge 1$, we let \[ Y_\cdot^n=(Y_i^n)_{i=0}^{\tilde{\tau}_n}:=(\tilde{X}_{\tilde{\tau}_n-i})_{i=0}^{\tilde{\tau}_n} \] be the reversed path of $(\tilde{X}_i)_{i=0}^{\tilde{\tau}_n}$. If $|z|$ is large enough, $d\ge 5$ and $n\ge 1$, then \begin{equation}\label{LVe8} E_{\hat{\mathbb P}}\check{\mathbb P}^{\tilde{X}_{\tilde{\tau}_n}+z} \big(X_\cdot\in A(Y_\cdot^n, z)\big)>C>0. \end{equation} \end{lemma} {\it Proof:}~ Let \[m_z:=\lfloor|z|^{1/2}\rfloor.\] Then \begin{align} &E_{\hat{\mathbb P}}\check{\mathbb P}^{\tilde{X}_{\tilde{\tau}_n}+z} \big(X_\cdot\notin A(Y_\cdot^n, z)\big)\nonumber\\ &\le E_{\hat{\mathbb P}}\check{\mathbb P}^{\tilde{X}_{\tilde{\tau}_n}+z}(\tau'_{m_z}\ge |z|/3)+ \hat{\mathbb{P}}(\tilde{\tau}_n-\tilde{\tau}_{n-m_z}\ge |z|/3)\label{LV*12}\\ &\quad +E_{\hat{\mathbb P}}\check{\mathbb P}^{\tilde{X}_{\tilde{\tau}_n}+z} (d(X_i,Y_\cdot^n)\le i^\delta\text{ for some } i>\tau'_{m_z})\label{LV*13}\\ &\quad +E_{\hat{\mathbb P}}\check{\mathbb P}^{\tilde{X}_{\tilde{\tau}_n}+z} (d(\tilde{X}_{\tilde{\tau}_n-j},X_\cdot)\le j^\delta\text{ for some }j>\tilde{\tau}_n-\tilde{\tau}_{n-m_z}).\label{LV*14} \end{align} We will first estimate (\ref{LV*12}). By the translation invariance of the environment measure, \[ \check{\mathbb P}^x(\tau'_{m_z}\ge |z|/3)=\check{\mathbb P}(\tau'_{m_z}\ge |z|/3) \text{ for any }x\in\mathbb{Z}^d. \] Hence \begin{equation}\label{LV2e4} E_{\hat{\mathbb P}}\check{\mathbb P}^{\tilde{X}_{\tilde{\tau}_n}+z}(\tau'_{m_z}\ge |z|/3) = \check{\mathbb P}(\tau'_{m_z}\ge |z|/3) \le \frac{3E_{\check{\mathbb P}}\tau'_{m_z}}{|z|} \stackrel{(\ref{LVetau})}{\le} C(E_{\check{\mathbb P}}\tau'_1)|z|^{-1/2}. \end{equation} Similarly, \begin{equation}\label{LV2e5} \hat{\mathbb P}(\tilde{\tau}_n-\tilde{\tau}_{n-m_z}\ge |z|/3) \stackrel{(\ref{LVprop2})}{\le} \exp{(e^{-cL})}\hat{\mathbb P}(\tau_{m_z}\ge |z|/3)\le C(E_{\hat{\mathbb P}}\tau_1) |z|^{-1/2}. \end{equation} To estimate (\ref{LV*13}) and (\ref{LV*14}), for $i\ge 1, n\ge j\ge 1$, we let \begin{align*} &Q'(i,x)=\check{\mathbb P}(x\text{ is visited in }[\tau'_{i-1},\tau'_i)),\\ &\tilde{Q}(j,x)=\hat{\mathbb P}(X_{\tau_n}+x \text{ is visited in}[\tau_{n-j},\tau_{n-j+1})). \end{align*} Note that by arguments that are similar to the proof of Theorem \ref{hke}, one can also obtain the heat kernel estimate (\ref{LVehke}) for $Q'(i,x)$ and $\tilde{Q}(j,x)$. For $l>0$, let $B(o,l)=\{x\in\mathbb{Z}^d: d(o,x)\le l\}$. Recall the definition of the $r$-boundary in Definition \ref{LVdef1}. By the translation invariance of the environment measure, \[ \check{\mathbb P}^y(X_i=y+z) =\check{\mathbb P}(X_i=z) \text{ for any }y,z\in\mathbb{Z}^d \text{and }i\in\mathbb{N}. \] Hence \begin{align*} &E_{\hat{\mathbb P}}\check{\mathbb P}^{\tilde{X}_{\tilde{\tau}_n}+z} (d(X_i,\tilde{X}_\cdot)\le i^\delta\text{ for some } i>\tau'_{m_z})\\ &\le \sum_{i\ge m_z}\sum_{y\in\partial_1 B(o,i^\delta)}\sum_x E_{\hat{\mathbb P}}\big[\check{\mathbb P}^{\tilde{X}_{\tilde{\tau}_n}+z} (\tilde{X}_{\tilde{\tau}_n}+z+x\text{ is visited in }[\tau'_i,\tau'_{i+1}))\\ &\qquad\qquad\qquad\qquad\qquad\times 1_{\tilde{X}_{\tilde{\tau}_n}+z+x+y\in Y_\cdot^n}\big]\\ &= \sum_{i\ge m_z}\sum_{y\in\partial_1 B(o,i^\delta)}\sum_x \check{\mathbb P}(x\text{ is visited in }[\tau'_i,\tau'_{i+1})) \hat{\mathbb P}(\tilde{X}_{\tilde{\tau}_n}+z+x+y\in Y_\cdot^n)\\ &=\sum_{i\ge m_z}\sum_{y\in\partial_1 B(o,i^\delta)}\sum_{j\le n}\sum_x Q'(i,x)\tilde{Q}(j,x+z+y). \end{align*} By the heat kernel estimates and H\"{o}lder's inequality, \begin{align*} \sum_{j\le n}\sum_x Q'(i,x)\tilde{Q}(j,x+z+y) &\le \sqrt{\sum_x Q'(i,x)^2}\sum_{j\le n}\sqrt{\sum_x \tilde{Q}(j,x+y)^2}\\ &\le C(E_{\check{\mathbb P}}\tau'_1)i^{-d/4} \sum_{j\le n}(E_{\hat{\mathbb P}}\tau_1)j^{-d/4}\\ &\stackrel{d\ge 5}{\le} Ci^{-d/4}E_{\check{\mathbb P}}\tau'_1 E_{\hat{\mathbb P}}\tau_1. \end{align*} Thus \begin{align}\label{LV2e6} &E_{\hat{\mathbb P}}\check{\mathbb P}^{\tilde{X}_{\tilde{\tau}_n}+z} (d(X_i,\tilde{X}_\cdot)\le i^\delta\text{ for some } i>\tau'_{m_z})\nonumber\\ &\le C\sum_{i\ge m_z}\sum_{y\in\partial_1 B(o,i^\delta)}i^{-d/4} E_{\check{\mathbb P}}\tau'_1 E_{\hat{\mathbb P}}\tau_1\nonumber\\ &\le C\sum_{i\ge m_z}i^{(d-1)\delta}i^{-d/4}E_{\check{\mathbb P}}\tau'_1 E_{\hat{\mathbb P}}\tau_1 \le C |z|^{-(d-4)/8} E_{\check{\mathbb P}}\tau'_1 E_{\hat{\mathbb P}}\tau_1, \end{align} where we used $d\ge 5$ and $\delta=\frac{d-4}{8(d-1)}$ in the last inequality. Similarly, we have \begin{align}\label{LV2e7} &E_{\hat{\mathbb P}}\check{\mathbb P}^{\tilde{X}_{\tilde{\tau}_n}+z} (d(\tilde{X}_{\tilde{\tau}_n-j},X_\cdot) \le j^\delta\text{ for some }j>\tilde{\tau}_n-\tilde{\tau}_{n-m_z})\nonumber\\ &\le C |z|^{-(d-4)/8}E_{\check{\mathbb P}}\tau'_1 E_{\hat{\mathbb P}}\tau_1. \end{align} Combining (\ref{LV2e4}), (\ref{LV2e5}), (\ref{LV2e6}) and (\ref{LV2e7}), we conclude that \begin{equation*} E_{\hat{\mathbb P}}\check{\mathbb P}^{\tilde{X}_{\tilde{\tau}_n}+z} \big(X_\cdot\in A(Y_\cdot^n, z)\big)>C>0, \end{equation*} if $|z|$ is large enough and $d\ge 5$.\qed\\ Let \[ T^o=\inf\{i\ge 0: X_i\cdot e_1<0\}. \] For every fixed $\omega\in\Omega$ and $P_{\omega,\epsilon}^o$-almost every $X_\cdot$, \[ P_{\omega,\theta^n\epsilon}^{X_n}(T^o=\infty)1_{T^o>n}=P_{\omega,\epsilon}^o (T^o=\infty|X_1,\ldots,X_n), \] and so by Levy's martingale convergence theorem, \[ \mathop{\rm lim}_{n\to\infty}P_{\omega,\theta^n\epsilon}^{X_n}(T^o=\infty)1_{T^o> n}= 1_{T^o=\infty}, \quad\text{$P_{\omega,\epsilon}^o$-almost surely}. \] Hence, for $(\omega, \epsilon,\tilde{X}_\cdot)$ sampled according to $\hat{\mathbb P}$, \[ \mathop{\rm lim}_{n\to\infty} P_{\omega,\theta^{\tilde{\tau}_n}\epsilon}^{\tilde{X}_{\tilde{\tau}_n}}(T^o=\infty)=1, \quad\text{$\hat{\mathbb P}$-almost surely}. \] It then follows by the dominated convergence theorem that \begin{equation}\label{LV3e7} \mathop{\rm lim}_{n\to\infty} E_{\hat{\mathbb P}} P_{\omega,\theta^{\tilde{\tau}_n}\epsilon}^{\tilde{X}_{\tilde{\tau}_n}}(T^o<\infty) =0. \end{equation} \begin{lemma}\label{LVl1} For any $z\in\mathbb{Z}^d$, \begin{equation}\label{LVe2} \mathop{\rm lim}_{n\to\infty} E_{\hat{\mathbb P}} P_{\omega,\theta^{\tilde{\tau}_n}\epsilon}^{\tilde{X}_{\tilde{\tau}_n}+z}(T^o<\infty) =0. \end{equation} \end{lemma} {\it Proof:}~ For $n>|z|$, obviously \[(\tilde{X}_{\tilde{\tau}_n}+z)\cdot e_1>0.\] This together with ellipticity yields \[ P_{\omega,\theta^{\tilde{\tau}_n}\epsilon}^{\tilde{X}_{\tilde{\tau}_n}}(T^o<\infty) \ge{(\frac{\kappa}{2})^{|z|}} P_{\omega,\theta^{\tilde{\tau}_n+|z|}\epsilon}^{\tilde{X}_{\tilde{\tau}_n}+z}(T^o<\infty). \] Hence using \eqref{LV3e7}, \[ \mathop{\rm lim}_{n\to\infty} E_{\hat{\mathbb P}} P_{\omega,\theta^{\tilde{\tau}_n+|z|}\epsilon}^{\tilde{X}_{\tilde{\tau}_n}+z}(T^o<\infty) =0. \] On the other hand, noting that $\{R>\tau_1\}=\{R=\infty\}$, \begin{align*} &E_{\hat{\mathbb P}} P_{\omega,\theta^{\tilde{\tau}_n+|z|}\epsilon}^{\tilde{X}_{\tilde{\tau}_n}+z}(T^o<\infty)\\ &= \sum_{m,x} E_{P\otimes Q}[P_{\omega,\theta^{m+|z|}\epsilon}^{x+z}(T^o<\infty) P_{\omega, \epsilon}^o (R>\tau_1,\tau_n=m,X_m=x)]/\mathbb{P}(R=\infty)\\ &=\sum_{m,x} E_{P\otimes Q}[P_{\omega,\theta^{m}\epsilon}^{x+z}(T^o<\infty) P_{\omega, \epsilon}^o (R>\tau_1,\tau_n=m,X_m=x)]/\mathbb{P}(R=\infty)\\ &= E_{\hat{\mathbb P}}P_{\omega,\theta^{\tilde{\tau}_n}\epsilon}^{\tilde{X}_{\tilde{\tau}_n}+z}(T^o<\infty), \end{align*} where we used the independence (under $Q$) of $P_{\omega, \theta^m\epsilon}^{x+z}(T^o<\infty)$ and $P_{\omega,\epsilon}^o(R>\tau_1,\tau_n=m, X_m=x)$ in the second to last equality. The conclusion follows. \qed\\ \noindent\textit{Proof of the uniqueness of the non-zero velocity when $d\ge 5$, as stated in Theorem \ref{LVthm2}:} If the two antipodal velocities are both non-zero, we assume that \[v_{e_1}\cdot v_{-e_1}>0.\] Sample $(\omega,\epsilon_\cdot,\tilde{X}_\cdot)$ according to $\hat{\mathbb P}$. Henceforth, we take $z=z_0$ such that (\ref{LVe8}) holds and \[z_0\cdot e_1<-L.\] We will prove Theorem \ref{LVthm2} by showing that \begin{equation}\label{LVcontradiction} E_{\hat{\mathbb P}} P_{\omega,\theta^{\tilde{\tau}_n}\epsilon}^{\tilde{X}_{\tilde{\tau}_n}+z_0}(T^o<\infty)>C \end{equation} for all $n>|z_0|$, which contradicts with (\ref{LVe2}). First, let $\mathcal{G}$ denote the set of finite paths $y_\cdot=(y_i)_{i=0}^M$ that satisfy $y_M=0, M<\infty$. Then \begin{align}\label{LVe11} & E_{\hat{\mathbb P}} P_{\omega,\theta^{\tilde{\tau}_n}\epsilon}^{\tilde{X}_{\tilde{\tau}_n}+z_0}(T^o<\infty)\\ &\ge E_{\hat{\mathbb P}} P_{\omega,\theta^{\tilde{\tau}_n}\epsilon}^{\tilde{X}_{\tilde{\tau}_n}+z_0} \big((X_i)_{i=0}^{T^o}\in A(Y_\cdot^n, z_0),T^o<\infty\big)\nonumber\\ &= \sum_{y_\cdot=(y_i)_{i=0}^M\in \mathcal{G}} E_{\hat{\mathbb P}} [P_{\omega,\theta^M\epsilon}^{y_0+z_0} \big((X_i)_{i=0}^{T^o}\in A(y_\cdot, z_0),T^o<\infty\big) 1_{Y_\cdot^n=y_\cdot}]\nonumber\\ &= \frac{1}{\mathbb{P}(R=\infty)} \sum_{y_\cdot\in \mathcal{G}} \sum_{\substack{N<\infty\\(x_i)_{i=0}^N\in A(y_\cdot, z_0)}} E_{P\otimes Q} [P_{\omega,\theta^M\epsilon}^{y_0+z_0} \big((X_i)_{i=0}^{T^o}=x_\cdot\big) P_{\omega,\epsilon}(Y_\cdot^n=y_\cdot)].\nonumber \end{align} By the definition of the regeneration times, for any finite path $y_\cdot=(y_i)_{i=0}^M$, there exists an event $G_{y_\cdot}$ such that $P_{\omega,\epsilon}(G_{y_\cdot})$ is $\sigma(\epsilon_{i,y_i}, \omega_{y_j}:0\le i\le M, 0\le j\le M-L)$-measurable and \[ \{Y_\cdot^n=y_\cdot\}=\{(\tilde{X}_i)_{i=0}^{\tilde{\tau}_n}=(y_{M-j})_{j=0}^M\} =G_{y_\cdot}\cap\{R\circ\theta_M=\infty\}. \] Hence, for and any $y_\cdot=(y_i)_{i=0}^M\in \mathcal{G}$ and $x_\cdot=(x_i)_{i=0}^N\in A(y_\cdot,z_0)$, $N<\infty$, \begin{align}\label{LV2e8} & E_{P\otimes Q} [P_{\omega,\theta^M\epsilon}^{y_0+z_0}\big((X_i)_{i=0}^{T^o}=x_\cdot, R_{-e_1}>N\big) P_{\omega,\epsilon}(Y_\cdot^n=y_\cdot)]\nonumber\\ &=E_P[ \bar{P}_\omega^{y_0+z_0}\big((X_i)_{i=0}^{T^o}=x_\cdot, R_{-e_1}>N\big) \bar{P}_\omega(G_{y_\cdot}) \bar{P}_\omega^{y_0}(R=\infty)]\nonumber\\ &\stackrel{(\ref{LVprop1})}{\ge} CE_P[ \bar{P}_\omega^{y_0+z_0}\big((X_i)_{i=0}^{T^o}=x_\cdot, R_{-e_1}>N\big) \bar{P}_\omega(G_{y_\cdot})] \bar{\mathbb P}(R=\infty). \end{align} where we used in the equality that $(\epsilon_{i,x})_{i\ge 0, x\in\mathbb{Z}^d}$ are iid and in the inequality the fact that \[ \bar{P}_\omega^{y_0+z_0}\big((X_i)_{i=0}^{T^o}=x_\cdot, R_{-e_1}>N\big) \bar{P}_\omega(G_{y_\cdot}) \] is $\sigma(\omega_v: v\cdot e_1\le y_0\cdot e_1-L)$-measurable (note that $z_0\cdot e_1<-L$). Further, by Lemma \ref{LVc2} and (\ref{LVe10}), we have \begin{align}\label{LV2e9} &E_P[ \bar{P}_\omega^{y_0+z_0}\big((X_i)_{i=0}^{T^o}=x_\cdot, R_{-e_1}>N\big) \bar{P}_\omega(G_{y_\cdot})]\nonumber\\ &\ge C\bar{\mathbb P}^{y_0+z_0}\big((X_i)_{i=0}^{T^o}=x_\cdot, R_{-e_1}>N\big) \bar{\mathbb P}(G_{y_\cdot}). \end{align} Note that \begin{align}\label{LVe12} \bar{\mathbb P}(G_{y_\cdot})\bar{\mathbb P}(R=\infty) &\stackrel{(\ref{LVprop1})}{\ge} CE_P [\bar{P}_\omega (G_{y_\cdot})\bar{P}_{\omega}^{y_0}(R=\infty)] \nonumber\\ &=C\bar{\mathbb P}(Y_\cdot^n=y_\cdot) \ge C\hat{\mathbb P}(Y_\cdot^n=y_\cdot). \end{align} Therefore, by (\ref{LVe11}), (\ref{LV2e8}) and (\ref{LV2e9}), \begin{align*} & E_{\hat{\mathbb P}} P_{\omega,\theta^{\tilde{\tau}_n}\epsilon}^{\tilde{X}_{\tilde{\tau}_n}+z_0}(T^o<\infty)\\ &\ge C\sum_{y_\cdot\in \mathcal{G}} \sum_{\substack{N<\infty\\(x_i)_{i=0}^N\in A(y_\cdot, z_0)}} \bar{\mathbb P}^{y_0+z_0}\big((X_i)_{i=0}^{T^o}=x_\cdot, R_{-e_1}>N\big) \bar{\mathbb P}(G_{y_\cdot})\bar{\mathbb P}(R=\infty)\\ &\stackrel{(\ref{LVe12})}{\ge} C\sum_{y_\cdot\in \mathcal{G}} \sum_{\substack{N<\infty\\(x_i)_{i=0}^N\in A(y_\cdot, z_0)}} \bar{\mathbb P}^{y_0+z_0}\big((X_i)_{i=0}^{T^o}=x_\cdot, R_{-e_1}>N\big) \hat{\mathbb P}(Y_\cdot^n=y_\cdot)\\ &\ge CE_{\hat{\mathbb P}} \check{\mathbb P}^{\tilde{X}_{\tilde{\tau}_n}+z_0} (X_\cdot\in A(Y_\cdot^n, z_0))\stackrel{\text{Lemma }\ref{LVl3}}{>}C. \end{align*} (\ref{LVcontradiction}) is proved.\qed \chapter{Bibliography} \chapter{Introduction} \label{intro_chapter} \section{An introduction to RWRE}\label{Section II} Let $\mathcal{M}=\mathcal{M}_1(V)$ be the space of all probability measures on $V=\{v\in\mathbb{Z}^d: |v|\le 1\}$, where $|\cdot|$ denotes the $l^2$-norm. We equip $\mathcal{M}$ with the weak topology on probability measures, which makes it into a Polish space, and equip $\Omega=\mathcal{M}^{\mathbb{Z}^d}$ with the induced Polish structure. Let $\mathcal{F}$ be the Borel $\sigma$-field of $\Omega$ and $P$ a probability measure on $\mathcal{F}$. A random \textit{environment} is an element $\omega =\{\omega(x, v)\}_{x\in{\mathbb{Z}^d}, v\in V}$ of $\Omega$. The random environment is called \emph{balanced} if \[P\{\omega(x, e_i)=\omega(x,-e_i) \mbox{ for all $i$ and all $x\in\mathbb{Z}^d$}\}=1,\] and \textit{elliptic} if $P\{\omega(x,e)>0 \mbox{ for all $|e|=1$ and all $x\in\mathbb{Z}^d$}\}=1$. We say that the random environment is \emph{uniformly elliptic} with ellipticity constant $\kappa$ if $P\{\omega(x,e)>\kappa \mbox{ for all $|e|=1$ and all $x\in\mathbb{Z}^d$}\}=1$. The random walk in the random environment $\omega\in\Omega$ (RWRE) started at $x$ is the Markov chain $\{X_n\}$ on $(\mathbb{Z}^d)^\mathbb{N}$, with state space $\mathbb{Z}^d$ and law $P_\omega^x$ specified by \begin{align*} &P_\omega^x\{X_0=x\}=1,\\ &P_\omega^x\{X_{n+1}=y+v | X_n=y\}=\omega(y, v), \quad v\in V. \end{align*} Let $\mathcal{G}$ be the $\sigma$-field generated by cylinder functions. The probability distribution $P_\omega^x$ on $((\mathbb{Z}^d)^\mathbb{N}, \mathcal{G})$ is called the \textit{quenched law}. Note that for each $G\in\mathcal{G}$, $P_\omega^x (G) : \Omega\to [0,1]$ is a $\mathcal{F}$-measurable function. The joint probability distribution $\mathbb{P}^x$ on $\mathcal{F}\times\mathcal{G}$: \[ \mathbb{P}^x (F\times G)=\int_{F} P_\omega^x (G)P(\, \mathrm{d}\omega), \qquad F\in\mathcal{F},\, G\in\mathcal{G}, \] is called the \textit{annealed} (or \textit{averaged}) law. Expectations with respect to $P_\omega^x$ and $\mathbb{P}^x$ are denoted by $E_\omega^x$ and $\mathbb{E}^x$, respectively. We also write $\mathbb{P}^o$ as $\mathbb{P}$, where $o=(0,\cdots, 0)$ is the origin. For $\omega\in\Omega$, set \[\omega_x=\big(\omega(x,e)\big)_{|e|=1}.\] Define the spatial shifts $\{\theta^y\}_{y\in\mathbb{Z}^d}$ on $\Omega$ by $(\theta^y\omega)_x=\omega_{x+y}$. We say that the random environment is \textit{ergodic} if the measure $P$ is ergodic with respect to the group of shifts $\{\theta^y\}$. A special case is when the probability vectors $(\omega_x)_{x\in\mathbb{Z}^d}$ are independent and identically distributed (\textit{iid}). Setting $\bar{\omega}(n)=\theta^{\mathrm{X}_n}\omega$, then the process $ \bar{\omega}(n) $ is a Markov chain under $ \mathbb{P}^{o} $ with state space $ \Omega $ and transition kernel \[ M(\omega',\, \mathrm{d}\omega)=\sum_{i=1}^{d}[\omega'(o,e_i)\delta_{\theta^{e_i}\omega'}+ \omega'(o,-e_i)\delta_{\theta^{-e_i}\omega'}]+\omega'(o,o)\delta_{\omega'}. \] $\big(\bar{\omega}(n)\big)_{n\in\mathbb{N}}$ is often referred as the ``\textit{environment viewed from the point of view of the particle}'' process. For $t\ge 0$, let \[ X_t=X_{\lfloor t\rfloor}+(t-\lfloor t\rfloor)(X_{\lfloor t\rfloor+1}-X_{\lfloor t\rfloor}). \] We say that the \textit{quenched invariance principle} of the RWRE holds if, for $P$-almost every $\omega\in\Omega$ and some deterministic vector $v\in\mathbb{R}^d$ (called the {\it limiting velocity}), the $P_\omega^o$ law of the path $\{(X_{tn}-tnv)/\sqrt{n}\}_{t\geq 0}$ converges weakly to a Brownian motion, as $n\to \infty$. For $\ell\in S^{d-1}$, we say that the RWRE is \emph{ballistic} in the direction $\ell$ if \[ \varliminf_{n\to\infty}\frac{X_n\cdot\ell}{n}>0,\quad\mathbb{P}\mbox{-a.s.} \] \section{Structure of the thesis} In this thesis, we study the diffusive and ballistic behaviors of random walks in random environment in $\mathbb{Z}^d, d\ge 2$. The organization of the thesis is as follows. Section~\ref{IOverview} gives an overview of the previous results in the study of the ballisticity, the central limit theorems (CLT), and the Einstein relation of RWRE. The three subsections in Section~\ref{Iresults} state the main results in this thesis and discuss the ideas of their proofs. Chapters \ref{LV chapter}, \ref{CLT chapter} and \ref{ER chapter} are devoted to the proofs of our three main results: In Chapter~\ref{LV chapter}, we consider the limiting velocity of random walks in strong-mixing random Gibbsian environments in $\mathbb{Z}^d, d\ge 2$. Based on regeneration arguments, we will first provide an alternative proof of Rassoul-Agha's conditional law of large numbers (CLLN) for mixing environment \cite{R-A3}. Then, using coupling techniques, we show that there is at most one nonzero limiting velocity in high dimensions ($d\ge 5$). Chapter~\ref{CLT chapter} proves the quenched invariance principles (Theorem~\ref{CLT1} and Theorem~\ref{CLT2}) for random walks in elliptic and balanced environments. We first prove an invariance principle (for $d\ge 2$) and the transience of the random walks when $d\ge 3$ (recurrence when $d=2$) in an ergodic environment which is not uniformly elliptic but satisfies certain moment condition. Then, using percolation arguments, we show that under (not necessarily uniform) ellipticity, the above results hold for random walks in iid balanced environments. Chapter~\ref{ER chapter} gives the proof of the Einstein relation in the context of random walks in a balanced uniformly elliptic iid random environment. Our approach combines a change of measure argument of Lebowitz and Rost \cite{Le} and the regeneration argument of Gantert, Mathieu and Piatnitski \cite{GMP}. The key step of our proof is the construction of a new regeneration structure. \section{Overview of previous results}\label{IOverview} \subsection{Ballisticity} The ballistic behavior of the RWRE in dimension $d\ge 2$ has been extensively studied. For random walks in iid random environment in dimension $d\ge 2$, the Kalikow's 0-1 law \cite{Ka81} states that for any direction $\ell\in S^{d-1}$, \[\mathbb{P}(A_\ell\cup A_{-\ell})\in\{0,1\}\] where $A_{\pm\ell}=\{\mathop{\rm lim}_{n\to\infty}X_n\cdot\ell=\pm\infty\}$. It is believed that for any direction $\ell$ and any $d\ge 2$, a stronger 0-1 law is true: \[ P(A_\ell)\in\{0,1\} \tag{0-1 Law}. \] When $d=2$, this $0$-$1$ law was proved by Zerner and Merkel \cite{ZM}. The question whether the 0-1 law holds for iid random environment in dimensions $d\ge 3$ is still open. (It is known that some strong mixing condition is necessary for the 0-1 law to hold, as the counterexample in \cite{BZZ} shows.) Much progress has been made in the study of the limiting velocity $\mathop{\rm lim}_{n\to\infty}X_n/n$ of random walks in iid environment, see \cite{ZO} for a survey. For one-dimensional RWRE, the law of large numbers (LLN) was proved in \cite{So}. For $d\ge 2$, a conditional law of large numbers (CLLN) was proved in \cite{SZ, Ze} (see \cite[Theorem 3.2.2]{ZO} for the full version). It states that $\mathbb{P}$-almost surely, for any direction $\ell$, \[ \mathop{\rm lim}_{n\to\infty}\frac{X_n\cdot\ell}{n}=v_+ 1_{A_\ell}-v_-1_{A_{-\ell}} \tag{CLLN} \] for some deterministic vectors $v_\ell$ and $v_{-\ell}$ (we set $v_\ell=o$ if $\mathbb{P}(A_\ell)=0$). This was achieved by considering the regenerations of the random walk path. Hence for $d\ge 2$, the 0-1 law would imply the LLN. Recall that when $d\ge 3$, the 0-1 law is one of the main open questions in the study of RWRE. Nevertheless, in high dimensions ($d\ge 5$), Berger \cite{Be} showed that the limiting velocity can take at most one non-zero value, i.e., \begin{equation}\label{Berger}v_\ell v_{-\ell}=0.\end{equation} It is of interests to consider environments whose law $P$ is not iid but rather ergodic (under possibly appropriate mixing conditions). Of special interest is the environment that is produced by a Gibbsian particle system (which we call the \textit{Gibbsian environment}) and satisfies Dobrushin-Shlosman's strong-mixing condition IIIc in \cite[page 378]{DS}, see \cite{R-A1,R-A2,CZ1,CZ2, R-A3} for related works. An important feature of this model is that the influence of the environments in remote locations decays exponentially as the distance grows. (We won't give the definitions of the Gibbsian environment and the strong-mixing condition in this thesis. For their definitions, we refer to \cite[pages 1454-1455]{R-A1}. We remark that our results only assume a mixing condition (G), which is defined in page \pageref{LVdef1}. It is known that (G) is a property of the strong-mixing Gibbsian environment, cf. \cite[Lemma 9]{R-A1}.) In \cite{R-A1}, assuming a ballisticity condition (Kalikow's condition) which implies that the event of escape in a direction has probability $1$, Rassoul-Agha proved the LLN for the strong-mixing Gibbsian environment, using the invariant measure of the ``environment viewed from the point of view of the particle" process $\big(\bar\omega(n)\big)$. In \cite{R-A3}, Rassoul-Agha also obtained the CLLN for the strong-mixing Gibbsian environment, under an analyticity condition (see Hypothesis (M) in \cite{R-A3}). Comets and Zeitouni proved the LLN for environments with a weaker cone-mixing assumption ($\mathcal{A}1$) in \cite{CZ1}, but under some conditions about ballisticity and the uniform integrability of the regeneration times (see ($\mathcal{A}5$) in \cite{CZ1}). \subsection{Central Limit Theorems} In recent years, there has been much interest in the study of invariance principles and transience/recurrence for random walks in random environments (on the $d$-dimensional lattice $\mathbb{Z}^d$) with non uniformly elliptic transitions probabilities. Much of this work has been in the context of reversible models, either for walks on percolation clusters or for the random conductance model, see \cite{Bar04,SS05,MR05,BB,MaP07,Ma08,BarDe10}. In those cases, the main issue is the transfer of annealed estimates (given e.g. in \cite{DFGW89} in great generality) to the quenched setting, and the control of the quenched mean displacement of the walk. On the other hand, in these models the reversibility of the walk provides for explicit expressions for certain invariant measures for the environment viewed from the point of view of the particle. The non-reversible setup has proved to provide many additional, and at this point insurmountable, challenges, even in the uniformly elliptic iid setup, see \cite{Zrev} for a recent account. In \cite{Sz3}, Sznitman shows that his condition (T') implies ballisticity and LLN and a directional annealed central limit theorem. The proof uses regeneration times and a renormalization argument and does not employ the process of the environment viewed from the point of view of particle. (We remark that weaker forms of the condition (T') exist, see \cite{Sz3, DR1, DR2, BDR}. Recently it was shown in \cite{BDR} that polynomial decay of some exit probabilities implies (T').) Further, it was shown by Berger and Zeitouni \cite{BZei} and Rassoul-Agha and Sepp\"{a}l\"{a}inen \cite{R-AS} that in the ballistic case, an annealed invariance principle is equivalent to a quenched invariance principle, under appropriate moment conditions on the regeneration times (these conditions are satisfied in all cases where a ballistic annealed CLT has been proved). When the walk is not ballistic, the regeneration structure employed in \cite{Sz2} is not available. Several classes of non-ballistic models were considered in the literature: balanced environment (see the definition in Section~\ref{Section II}), environment whose sufficiently high-dimensional projection is a simple random walk \cite{BSZ}, and isotropic environment which is a small perturbation of the simple random walk \cite{BK, BolZei, SZei}. Historically, the first to be considered was the balanced environment, first investigated by Lawler \cite{La}, which we describe next as a good part of the thesis deals with that environment: \begin{theorem}[\cite{La},\cite{ZO}]\label{LaThm} Assume the random environment is ergodic, balanced and uniformly elliptic. Then $P$-almost surely, the $P_\omega$ law of the rescaled path $\lambda X_{\cdot/\lambda^2}$ converges weakly to a Brownian motion on $\mathbb{R}^d$ with a non-degenerate diagonal covariance matrix. Moreover, the RWRE is recurrent for $d=2$ and transient for $d\ge 3$, $P$-almost surely. \end{theorem} In this case, a-priori estimates of the Alexandrov-Bakelman-Pucci type give enough control that allows one to prove the existence of invariant measures (for the environment viewed from the point of view of the particle), and the fact that the walk is a (quenched) martingale together with ergodic arguments yield the invariance principle (obviously, control of the quenched mean displacement, which vanishes, is automatic). The establishment of recurrence (for $d=2$) and transience (for $d\geq 3$) requires some additional arguments, due to Kesten and Lawler, respectively, see \cite{ZO} for details. \subsection{Einstein relation} In 1905, Einstein \cite[pp. 1-18]{Einstein} investigated the movement of suspended particles in a liquid under the influence of an external force. He established the following linear relation between the diffusion constant $D$ and the \textit{mobility} $\mu$: \[ D\sim T\mu, \] where $T$ is the absolute temperature, and $\mu$ is defined as the limiting ratio between the velocity (under the external force) and the force, as the force goes to zero. More precisely, the Einstein relation (ER) describes the relation between the response of a system to a perturbation and its diffusivity at equilibrium. It states that the derivative of the velocity (with respect to the strength of the perturbation) equals the diffusivity: \[ \mathop{\rm lim}_{\lambda\to 0}\mathop{\rm lim}_{t\to\infty}\frac{E_\lambda X_t/t}{\lambda}=D, \tag{ER} \] where $(X_t)_{t\ge 0}\in (\mathbb{R}^d)^{\mathbb{R}_+}$ denotes the random motion of the particle, $\lambda$ is the size of the perturbation, $D$ is the diffusion constant of the equilibrium state, and $E_\lambda$ is the annealed measure of the perturbed media. General derivations of this principle assume reversibility. Recently, there has been much interest in studying the Einstein relation for reversible motions in random media, see \cite{Le,KO,GMP,BHOZ}. In \cite{Le}, Lebowitz and Rost proved a weak form of the Einstein relation for a wide class of random motions in random media: \[ \mathop{\rm lim}_{\lambda\to 0}E_\lambda \frac{X_{t/\lambda^2}}{t/\lambda}=D \quad \forall t>0. \] In \cite{KO}, the ER is verified for random walks in random conductance, where the conductance is only allowed to take two values. The approach of \cite{KO} is an adaption of the perturbation argument and transience estimates in \cite{Loulakis}. For random walks on Galton-Watson trees, the ER is proved by \cite{BHOZ}. Their approach uses recursions due to the tree structure and renewal arguments. Recently, Gantert, Mathieu and Piatnitski \cite{GMP} established the ER for random walks in random potential, by combining the argument in \cite{Le} with good moment estimates of the regeneration times. The Einstein relation for random motions in the non-reversible zero speed set-up, e.g., random walks in balanced random environments (RWBRE), is a challenging problem. (In general one expects correction terms in (ER) due to the non-reversibility of the walk.) \section{Our results}\label{Iresults} In this section we will state the main results in the thesis and explain the ideas of their proofs. The actual proofs will be presented in the following chapters. Our contributions are in three directions: CLLN and regeneration structures for RWRE in Gibbsian environments, quenched invariance principles for balanced elliptic (but non uniformly elliptic) environments, and ER for balanced iid uniformly elliptic environments. \subsection{Limiting velocity for mixing random environment}\label{ILV} Recall first the definition of an $r$-Markov environment (see \cite{CZ2}). \begin{definition}\label{LVdef1} For $r\ge 1$, let $\partial_r V=\{x\in\mathbb{Z}^d\setminus V: d(x, V)\le r\}$ be the $r$-boundary of $V\subset\mathbb{Z}^d$. A random environment $(P,\Omega)$ on $\mathbb{Z}^d$ is called $r$-Markov if for any finite $V\subset\mathbb{Z}^d$, \[ P\big((\omega_x)_{x\in V}\in \cdot|\mathcal{F}_{V^c}\big) =P\big((\omega_x)_{x\in V}\in \cdot|\mathcal{F}_{\partial_r V}\big), \text{ $P$-a.s.,} \] where $d(\cdot,\cdot)$ denotes the $l^1$-distance and $\mathcal{F}_{\Lambda}:=\sigma(\omega_x:x\in\Lambda)$. \end{definition} We say that an $r$-Markov environment $P$ {\it satisfies condition (G)} if there exist constants $\gamma , C<\infty$ such that for all finite subsets $\Delta\subset V\subset\mathbb{Z}^d$ with $d(\Delta,V^c)\ge r$, and $A\subset V^c$, \[ \frac{\, \mathrm{d} P\big((\omega_x)_{x\in\Delta}\in\cdot|\eta\big)} { \, \mathrm{d} P\big((\omega_x)_{x\in\Delta}\in\cdot|\eta'\big)} \le \exp{(C\sum_{x\in A,y\in\Delta}e^{-\gamma d(x,y)})}\tag{G} \] for $P$-almost all pairs of configurations $\eta,\eta'\in\mathcal{M}^{V^c}$ which agree on $V^c\setminus A$. Here \[ P\big((\omega_x)_{x\in\Delta}\in\cdot|\eta\big) :=P\big((\omega_x)_{x\in\Delta}\in \cdot|\mathcal{F}_{V^c}\big)\big|_{(\omega_x)_{x\in V^c}=\eta}. \] We remark that $r$ and $\gamma$ are used as parameters of the environment throughout the article. Recall that by Lemma 9 in \cite{R-A1}, the strong-mixing Gibbsian environment satisfies (G). Obviously, every finite-range dependent environment also satisfies (G). Our main theorem concerning the mixing environments is: \begin{theorem}\label{LVthm2} Assume that $P$ is uniformly elliptic and satisfies \emph{(G)}. Then there exist two deterministic constants $v_+, v_-\ge 0$ and a vector $\ell$ such that \begin{equation}\label{ICLLN} \mathop{\rm lim}_{n\to\infty} \frac{X_n}{n}=v_+\ell 1_{A_\ell}-v_-\ell 1_{A_{-\ell}}, \end{equation} and $v_+=v_-=0$ if $\mathbb{P}(A_\ell\cup A_{-\ell})<1$. Moreover, if $d\ge 5$, then there is at most one non-zero velocity. That is, \begin{equation}\label{Iunique} v_+ v_-=0. \end{equation} \end{theorem} We remark here that for the finite-range dependent case, the CLLN is proved in \cite{ZO}. \eqref{ICLLN} is a minor extension of Rassoul-Agha's CLLN in \cite{R-A3}. He assumes slightly more than strong-mixing, which in turn is slightly stronger than our condition (G). Our proof is very different from the proof in \cite{R-A3} , which is based on a large deviation principle in \cite{R-A2}. The main contribution of our proof of \eqref{ICLLN} is a new definition of the regeneration structure, which enables us to divide a random path in the mixing environment into ``almost iid" parts. With this regeneration structure, we will use the ``$\epsilon$-coins" introduced in \cite{CZ1} and coupling arguments to prove the CLLN. This regeneration structure will also be used in the proof of \eqref{Iunique}. Display \eqref{Iunique} is an extension of Berger's result \eqref{Berger} from the iid case to our case (G), which includes the strong-mixing case. In \cite{Be}, assuming that $\mathbb{P}(A_\ell)>0$ for a direction $\ell$, Berger coupled the iid environment $\omega$ with a transient (in the direction $\ell$) environment $\tilde\omega$ and a ``backward path", such that $\tilde\omega$ and $\omega$ coincide in the locations off the path. Using heat kernel estimates for random walks with iid increments, he showed that if $v_\ell v_{-\ell}>0$ and $d\ge 5$, then with positive probability, the random walks in $\tilde\omega$ is transient to the $-\ell$ direction without intersecting the backward path, which contradicts $\tilde{\omega}$ being transient in the direction $\ell$. The difficulties in applying this argument to mixing environments are that the regeneration slabs are not iid, and that unlike the iid case, the environments visited by two disjoint paths are not independent. To overcome these difficulties, we will construct an environment (along with a path) that is ``very transient" in $\ell$, and show that the ballistic walks in the opposite direction $-\ell$ will move further and further away from the given path (see Figure \ref{LVfig:1} in Section \ref{secunique}). The key ingredient here is a heat kernel estimate, which we will obtain in Section \ref{sechke} using coupling arguments. \subsection{Invariance principle for RWBRE} As mentioned above, Lawler \cite{La} proved the invariance principle under the uniform ellipticity assumption. We explore the extent to which the uniform ellipticity assumption can be dropped. Surprisingly, in the iid case, we can show that no assumptions of uniform ellipticity are needed at all. Let \begin{equation} \label{CLTepsdef} \varepsilon(x)=\varepsilon_{\omega}(x):= [\prod_{i=1}^{d}\omega(x,e_i)]^{\frac{1}{d}}. \end{equation} Our first main result is that if $\mathrm{E}\varepsilon(o)^{-p}< \infty$ for some $p>d$, then the quenched invariance principle holds and moreover, the RWRE is transient $P$-almost surely if $d\geq 3$. (Recurrence for $d=2$ under the condition $E\varepsilon(0)^{-p}<\infty$ follows from the quenched invariance principle and ergodicity by an unpublished argument of Kesten detailed in \cite[Page 281]{ZO}. Note that this argument cannot be used to prove transience in dimensions $d\geq 3$, even given an invariance principle, since in higher dimensions the invariance principle does not give useful information on the range of the random walk; the behavior of the range is a crucial element in Kesten's argument.) \begin{theorem}\label{CLT1} Assume that the random environment is ergodic, elliptic and balanced. \begin{enumerate} \item[(i)] If $E\varepsilon(o)^{-p}< \infty$ for some $p>d\ge 2$, then the quenched invariance principle holds with a non-degenerate diagonal limiting covariance matrix. \item[(ii)] If $E[(1-\omega(o,o))/\varepsilon(o)]^q< \infty$ for some $q>2$ and $d\ge 3$, then the RWRE is transient $P$-almost surely. \end{enumerate} \end{theorem} \noindent That some integrability condition on the tail of $\varepsilon(o)$ is needed for part (i) to hold is made clear by the (non-Gaussian) scaling limits of random walks in Bouchaud's trap model, see \cite{Bou,BAC}. In fact, it follows from that example that Theorem \ref{CLT1}(i), or even an annealed version of the CLT, cannot hold in general with $p<1$. The proof of Theorem \ref{CLT1} is based on a sharpening of the arguments in \cite{La,Sz1,ZO}; in particular, refined versions of the maximum principle for walks in balanced environments (Theorem \ref{CMP}) and of a mean value inequality (Theorem \ref{Cmvi}) play a crucial role. When the environment is iid and elliptic, our second main result is that if $|X_{n+1}-X_n|=1$ a.s., then the quenched invariance principle holds. Moreover, the RWRE is $P$-almost surely transient when $d\ge 3$. The proofs combine percolation arguments with Theorem \ref{CLT1}. \begin{theorem}\label{CLT2} Assume that the random environment is iid, elliptic and balanced. \begin{enumerate} \item[(i)] If $P\{\max_{|e|=1}\omega(o,e)\ge \xi_0\}$=1 for some positive constant $\xi_0$, then the quenched invariance principle holds with a non-degenerate limiting covariance. \item[(ii)] When $d\ge 3$, the RWRE is transient $P$-almost surely. \end{enumerate} \end{theorem} Because the transience or recurrence of the random walks does not change if one considers the walk restricted to its jump times, one concludes, using Kesten's argument and the invariance principle, comparing with Theorem \ref{CLT1}, that for $d=2$, a random walk in a balanced elliptic iid random environment is recurrent $P$-a.s. Our proof of the invariance principles, like that of \cite{La}, is based on the approach of the ``environment viewed from the point of view of the particle". Since $\{X_n\}$ is a (quenched) martingale, standard arguments (see the proof of Theorem 6.2 in \cite{BB}) show that the quenched invariance principle holds whenever an invariant measure $Q\sim P$ of $\{\bar{\omega}(n)\}$ exists. The approach of Lawler \cite{La}, which is a discrete version of the argument of Papanicolaou and Varadhan \cite{PV}, is to construct such a measure as the limit of invariant measures of periodized environments. We will follow this strategy using, as in \cite{Sz1,ZO}, variants of \cite{KT} to derive estimates on solutions of linear elliptic difference equations. In the iid setup of Theorem~\ref{CLT2}, percolation estimates are used to control pockets of the environment where those estimates are not strong enough. For the proof of the transience in the ergodic case, we use a mean value inequality and follow \cite{ZO}. To prove the transience in the iid case, we employ percolation arguments together with a new maximum principle (Theorem \ref{Cmp2}) for walks with (possibly) big jumps. \begin{remark} Recently, Berger and Deuschel \cite{BD} have generalized our ideas and extended the quenched invariance principle to the general non-elliptic case where the environment is only required to be iid and “genuinely $d$-dimensional”. \end{remark} \subsection{Einstein relation for RWBRE}\label{IER} In this subsection we will present the Einstein relation for random walks in uniformly elliptic balanced iid random environment. Recall that by Theorem~\ref{LaThm}, for $P$-almost every $\omega$, $(\lambda X_{t/\lambda^2})_{t\ge 0}$ converges weakly (as $\lambda\to 0$) to a Brownian motion with a non-degenerate covariance matrix, which we denote by $\bm{D}$. For $\lambda\in (0,1)$ and a fixed direction \[ \ell=(\ell_1,\ldots,\ell_d)\in S^{d-1}, \] define the perturbed environment $\omega^\lambda$ of $\omega\in\Omega$ by \[ \omega^\lambda(x,e)=(1+\lambda\ell\cdot e)\omega(x,e). \] Since $\omega^\lambda$ satisfies Kalikow's condition (see (0.7) in \cite{SZ}), it follows from \cite[Theorem 2.3]{SZ} that there exists a deterministic constant $v_\lambda\in\mathbb{R}^d$ such that \[ \mathop{\rm lim}_{t\to\infty} \frac{X_t}{t}=v_\lambda, \quad \text{ $P\otimes P_{\omega^\lambda}^o$-almost surely}. \] Our main result is the following mobility-diffusivity relation: \begin{equation}\label{Einstein relation} \mathop{\rm lim}_{\lambda\to 0}\frac{v_\lambda}{\lambda}=D_\ell, \end{equation} where \[ D_\ell:=\bm{D}\ell=(2E_Q\omega(o,e_i)\ell_i)_{1\le i\le d}\in \mathbb{R}^d. \] Our proof of the Einstein relation \eqref{Einstein relation} consists of proving the following two theorems: \begin{theorem}\label{ER1} Assume that the environment $P$ is iid, balanced and uniformly elliptic. Then for $P$-almost every $\omega$ and for any $t\ge 1$, \begin{equation*} \mathop{\rm lim}_{\lambda\to 0} E_{\omega^\lambda}\frac{X_{t/\lambda^2}}{t/\lambda}=D_\ell. \end{equation*} \end{theorem} \begin{theorem}\label{ER2} Assume that the environment $P$ is iid, balanced and uniformly elliptic. Then for all sufficiently small $\lambda\in (0,1)$ and any $t\ge 1$, \[ \left|E_PE_{\omega^\lambda}\frac{X_{t/\lambda^2}}{t/\lambda}-\frac{v_\lambda}{\lambda}\right| \le \frac{C}{t^{1/5}}. \] \end{theorem} Our proof of Theorem \ref{ER1} is an adaption of the argument of Lebowitz and Rost \cite{Le} (see also \cite[Proposition 3.1]{GMP}) to the discrete setting. Namely, using a change of measure argument, we will show that the scaled process $\lambda X_{t/\lambda^2}$ converges (under the law $P_{\omega^\lambda}$) to a Brownian motion with drift $tD_\ell$, which yields Theorem \ref{ER1}. For the proof of Theorem \ref{ER2}, we want to follow the strategy of Gantert, Mathieu and Piatnitski \cite{GMP}. Arguments in the proof of \cite[Proposition 5.1]{GMP} show that if we can construct a sequence of random times $(\tau_n)_{n\in\mathbb{N}}$ (called the {\it regeneration times}) that divides the random path into iid (under the annealed measure) pieces, then good moment estimates of the regeneration times yield Theorem \ref{ER2}. In the construction of the regeneration times in \cite{GMP}, a heat kernel estimate \cite[Lemma 5.2]{GMP} for reversible diffusions is crucially employed. However, due to the lack of reversibility, we don't have a good heat kernel estimate for RWRE. In this thesis, we construct the regeneration times differently, so that they divide the random path into ``almost iid" parts. Moreover, our regeneration times have good moment bounds, which lead to a proof of Theorem \ref{ER2}. The key ingredients in our construction are Kuo and Trudinger's \cite{KT} Harnack inequality for discrete harmonic functions and the ``$\epsilon$-coins" trick introduced by Comets and Zeitouni \cite{CZ1}.
\section{} \tableofcontents \newpage \section{Introduction} \label{Introduction} The origin of the four-dimensionality of the present universe is one of the most fundamental problems in gravitational physics. One possible explanation is that only the four-dimensional universe is stable in some physical sense and therefore chosen at the moment of its creation. Another possibility is that after the creation of a higher-dimensional universe, the extra spatial dimensions other than our perceived three-dimensional space are compactified by some mechanism. Einstein's general theory of relativity allows us to study this fundamental problem by setting the number, $n$, of spacetime dimensions as a tunable parameter. It is therefore possible to look for critical values of $n$ beyond which spacetime properties change drastically. Such analyses can give us valuable insights about the origin of our four dimensional universe. In this context, it is reasonable to focus on black holes because they are fundamental objects that encode many key features of the gravitational interaction. In four dimensions, asymptotically flat stationary black holes are characterized by a small number of parameters such as mass or angular momentum, somewhat analogous to atomic properties in chemistry. This is a consequence of the black-hole uniqueness theorem asserting that the Kerr-Newman black hole is the unique asymptotically flat stationary and rotating black hole with a connected horizon in the Einstein-Maxwell system. (See~\cite{BHuniqueness} for review.) It is important to note that this uniqueness theorem is not valid in higher dimensions~\cite{er2008}, as explicitly shown in five dimensions, for example, by the existence of two distinct asymptotically flat black objects with the same mass and angular momentum but with different horizon topology~\cite{myersperry1986,er2002}. This fact already demonstrates one special feature of four-dimensional spacetime. $n=4$ is further singled out as a critical value in the framework of general relativity because asymptotically flat vacuum black holes do not exist in $n<4$ dimensions~\cite{ida2000}. When studying higher dimensions, it must be remembered that general relativity is not the only natural extension of four-dimensional Einstein gravity. General relativity is a quasi-linear second-order theory, which ensures the well-definedness of the initial value problem and the absence of ghosts. In 1971 Lovelock showed that in higher dimensions general relativity is just a special case of the most general class of theories satisfying these property. These more general theories are collectively called Lovelock gravity~\cite{Lovelock}. The Lovelock Lagrangian consists of a sum of the dimensionally extended Euler densities in which the cosmological constant and the Einstein-Hilbert terms appear as the zeroth- and the first-order terms, respectively. Just as the Einstein tensor trivially vanishes in two-dimensional spacetime, the second-order Lovelock Lagrangian (called the Gauss-Bonnet term) becomes purely topological in four spacetime dimensions and does not contribute to the field equations~\cite{lanczos}. As a consequence, Lovelock gravity reduces to general relativity with a cosmological constant in four dimensions. Motivation for studying Lovelock gravity is also provided by the fact that string/M-theory~\cite{string} requires the existence of extra spatial dimensions whereas quantum theory suggests the need to add higher-curvature terms. Moreover, string theoretic arguments~\cite{string lovelock} suggest that quadratic Lovelock gravity appears in the low-energy limit for strings propagating in curved spacetime. For the above reasons, Lovelock gravity has been extensively investigated, with emphasis on the similarities to and difference from general relativity. (See~\cite{lovelockreview} for review.) However, in comparison with its classical aspects, the quantum theory is poorly understood at present, despite the fact that deep issues such as the origin of the four-dimensional universe can only be answered in the quantum context. In four-dimensional general relativity, Kucha\v{r} presented an elegant geometrical framework for studying the physical phase space and quantization of spherically symmetric vacuum black holes~\cite{kuchar94}. Using Kucha\v{r} geometrodynamics as a foundation, Louko and M\"akel\"a were able present a rigorous quantization of the Schwarzschild black hole spacetime, including a construction of all self-adjoint extensions of the Hamiltonian and derivation of the semi-classical area spectrum~\cite{Louko1996}. They found that the area/entropy spectrum was equally spaced in the semi-classical limit, in agreement with early speculations of Bekenstein and Mukhanov based on the thermodynamic properties of black holes~\cite{Bekenstein}. An equally spaced area spectrum was also obtained for Schwarzschild black holes using a variety of different techniques~\cite{area spectrum}. This result is perhaps not surprising because there is only one length scale in the system, namely the Planck length. In contrast, there is more than one length scale in quantum Lovelock gravity because the coupling constants to each order of the Lovelock Lagrangian are dimensionful. As a consequence, the entropy of Lovelock black holes is no longer equal to 1/4 the area~\cite{whitt1988}. It is therefore of great interest to see what quantum spectrum emerges for the both the area and entropy. It is important to mention two key features of general relativity that were crucial to Kucha\v{r}'s analysis. First the existence of Birkhoff's theorem implies that the reduced phase space is finite dimensional (two dimensional in the case of Schwarzschild black holes). Secondly, Kucha\v{r} identified the Misner-Sharp mass~\cite{ms1964} in spherically symmetric vacuum spacetime and its conjugate momentum, the Schwarzschild time separation at infinity, as the physical phase space variables. In more general situations (i.e. non-vacuum), the Misner-Sharp mass is known as the best quasi-local mass in spherically symmetric spacetime. It satisfies the requisite monotonic and positivity properties and converges to the Arnowitt-Deser-Misner (ADM) mass at spacelike infinity in asymptotically flat spacetime~\cite{hayward1996}. Fortunately, Lovelock gravity also possesses both these key features. Bikrhoff's theorem in Lovelock gravity asserts that the spherically symmetric vacuum solution is uniquely determined under certain conditions~\cite{zegers2005}. The corresponding Schwarzschild-Tangherlini-type vacuum solution was obtained by Zegers~\cite{zegers2005}. In addition, a natural counterpart to the Misner-Sharp mass has been defined in Lovelock gravity~\cite{mwr2011,HM08}. The purpose of the present paper is to provide a framework to study quantum aspects of spherically symmetric black holes in Lovelock gravity. We provide a comprehensive analysis that goes far beyond the initial presentation of our results in~\cite{GTMletter}. In particular, we use the geometrodynamical formulation of Kucha\v{r} to do a complete Hamiltonian analysis, including derivation of the super-Hamiltonian and super-momentum constraints and verification of suitable boundary conditions for asymptotically flat black holes. Our analysis leads to a fully reduced Hamiltonian that is just as simple as that of Kucha\v{r}. As a specific application, we also derive the fully reduced equations of motion in flat slice coordinates for the collapse of a charged scalar field, including Lovelock gravitational as well as electromagnetic self-interactions. We note that the Hamiltonian analysis for full Lovelock gravity was first considered by Teitelboim and Zanelli~\cite{TZ}. Their result was rather formal in that an explicit parametrization of the phase space was not provided. (See also~\cite{df2012,st2008}.) For the case of spherical symmetry, the geometrodynamics~\cite{kuchar94} of five-dimensional Einstein-Gauss-Bonnet (i.e. quadratic Lovelock) gravity was worked out by Louko {\it et al}~\cite{JL97}, while the Hamiltonian analysis of higher-dimensional Gauss-Bonnet gravity coupled to matter was recently done in~\cite{TLKM}. Our analysis was done for generic Lovelock gravity in arbitrary dimensions. In the following section, we present our system, including action and spherically symmetric solutions. Section~\ref{Dim red action} derives a dimensionally reduced equivalent two-dimensional action that is the starting point of our analysis. It also reviews the geometrodynamics of Kucha\v{r} in a general dynamical setting. In Section~\ref{Can formalism gr}, we perform the Hamiltonian analysis for general relativity, in terms of both the standard ADM and geometrodynamical variables. The generalization to Lovelock gravity is presented in Section~\ref{Can form Lovelock}. The contributions of matter fields are discussed in Section~\ref{matter}, while concluding remarks and discussions appear in Section~\ref{Conclusions}. Detailed derivations and analysis of the boundary conditions are deferred to Appendices. Our basic notation follows~\cite{wald}. The convention for the Riemann curvature tensor is $[\nabla _\rho ,\nabla_\sigma]V^\mu ={{\cal R}^\mu }_{\nu\rho\sigma}V^\nu$ and ${\cal R}_{\mu \nu }={{\cal R}^\rho }_{\mu \rho \nu }$. The Minkowski metric is taken as diag$(-,+,\cdots,+)$, and Greek indices run over all spacetime indices. We adopt the units in which only the $n$-dimensional gravitational constant $G_n$ is retained. \section{Preliminaries} \label{Preliminaries} \subsection{Symmetric spacetimes in Lovelock gravity} The action of the gravitational system is written as \begin{align} I=I_{{\cal M}}+I_{\partial{\cal M}},\label{action} \end{align} where $I_{{\cal M}}$ is the dynamical term and $I_{\partial{\cal M}}$ is the boundary term. In general relativity, $I_{{\cal M}}$ is the Einstein-Hilbert action and $I_{\partial{\cal M}}$ is the Gibbons-Hawking-York boundary term. (See~\cite{olea2007} for the boundary term in general Lovelock gravity.) In the present paper, we consider Lovelock gravity in $n (\geq 4)$-dimensional spacetime, of which the dynamical term in the action is given by \begin{align} \label{action2} I_{{\cal M}}=&\frac{1}{2\kappa_n^2}\int \D ^nx\sqrt{-g}\sum_{p=0}^{[n/2]}\alpha_{(p)}{\ma L}_{(p)}+I_{\rm matter},\\ {\ma L}_{(p)}:=&\frac{1}{2^p}\delta^{\mu_1\cdots \mu_p\nu_1\cdots \nu_p}_{\rho_1\cdots \rho_p\sigma_1\cdots \sigma_p}{\cal R}_{\mu_1\nu_1}^{\phantom{\mu_1}\phantom{\nu_1}\rho_1\sigma_1}\cdots {\cal R}_{\mu_p\nu_p}^{\phantom{\mu_p}\phantom{\nu_p}\rho_p\sigma_p}, \end{align} where $\kappa_n := \sqrt{8\pi G_n}$. Our notation basically follows~\cite{mwr2011}. $\alpha_{(p)}$ is the coupling constant for the $p$th-order Lovelock Lagrangian with dimension $({\rm length})^{2(p-1)}$ and we assume $\kappa_n^2>0$ without any loss of generality. The $\delta$ symbol denotes a totally anti-symmetric product of Kronecker deltas, normalized to take values $0$ and $\pm 1$, defined by \begin{align} \delta^{\mu_1\cdots \mu_p}_{\rho_1\cdots \rho_p}:=&p!\delta^{\mu_1}_{[\rho_1}\cdots \delta^{\mu_p}_{\rho_p]}. \end{align} The gravitational equation following from this action is given by \begin{align} {\ma G}_{\mu\nu}=\kappa_n^2 {T}_{\mu\nu}, \label{beqL} \end{align} where ${T}_{\mu\nu}$ is the energy-momentum tensor for matter fields obtained from $I_{\rm matter}$ and \begin{align} {\ma G}_{\mu\nu} :=& \sum_{p=0}^{[n/2]}\alpha_{{(p)}}{G}^{(p)}_{\mu\nu}, \label{generalG}\\ {G}^{\mu(p)}_{~~\nu}:=& -\frac{1}{2^{p+1}}\delta^{\mu\eta_1\cdots \eta_p\zeta_1\cdots \zeta_p}_{\nu\rho_1\cdots \rho_p\sigma_1\cdots \sigma_p}{\cal R}_{\eta_1\zeta_1}^{\phantom{\eta_1}\phantom{\zeta_1}\rho_1\sigma_1}\cdots {\cal R}_{\eta_p\zeta_p}^{\phantom{\eta_p}\phantom{\zeta_p}\rho_p\sigma_p}. \end{align} The tensor ${G}^{(p)}_{\mu\nu}$ obtained from ${\ma L}_{(p)}$ contains up to the second derivatives of the metric and ${G}^{(p)}_{\mu\nu}\equiv 0$ is satisfied for $p\ge [(n+1)/2]$. In the present paper, we consider the $n(\ge 4)$-dimensional warped product spacetime $({\cal M}^n,g_{\mu\nu}) \approx ({M}^2,g_{AB})\times ({K}^{n-2},\gamma_{ab})$ with the general metric \begin{eqnarray} g_{\mu\nu}(x)dx^\mu dx^\nu=g_{AB}({\bar y})d{\bar y}^A d{\bar y}^B+R({\bar y})^2\gamma_{ab}(z)dz^adz^b, \label{eq:structure} \end{eqnarray} where $g_{AB}$ is an arbitrary Lorentz metric on $({M}^2,g_{AB})$ and $R({\bar y})$ is a scalar function on $({M}^2,g_{AB})$. $\gamma_{ab}$ is the metric on the $(n-2)$-dimensional maximally symmetric space $({K}^{n-2},\gamma_{ab})$ with its sectional curvature $k=1,0,-1$. We note that the results in the present paper are valid for $k=0$ with $p=0$ by setting $k^p=1$. We introduce the covariant derivatives on spacetime $({\ma M}^n,g_{\mu\nu})$, the subspacetime $({M}^2,g_{AB})$ and the maximally symmetric space $({K}^{n-2},\gamma_{ab})$ with \begin{eqnarray} \nabla_\rho g_{\mu\nu}=0,\qquad D_F g_{AB}=0,\qquad {\bar D}_fg_{ab}=0. \end{eqnarray} The most general energy-momentum tensor $T_{\mu\nu}$ compatible with this spacetime symmetry governed by Lovelock equations is given by \begin{align} T_{\mu\nu}\D x^\mu \D x^\nu =T_{AB}({\bar y})\D {\bar y}^A\D {\bar y}^B+p({\bar y})R^2 \gamma_{ab}\D z^a\D z^b, \end{align} where $T_{AB}({\bar y})$ and $p({\bar y})$ are a symmetric two-tensor and a scalar on $(M^2, g_{AB})$, respectively. The generalized Misner-Sharp mass in Lovelock gravity is defined by \begin{align} M :=& \frac{(n-2)V_{n-2}^{(k)}}{2\kappa_n^2}\sum_{p=0}^{[n/2]}{\tilde \alpha}_{(p)}R^{n-1-2p}[k-(DR)^2]^p,\label{qlm-L}\\ {\tilde \alpha}_{(p)}:=&\frac{(n-3)!\alpha_{(p)}}{(n-1-2p)!}, \label{alphatil} \end{align} where $(DR)^2:=(D_A R)(D^A R)$~\cite{mwr2011}. The constant $V_{n-2}^{(k)}$ represents the volume of $({K}^{n-2},\gamma_{ab})$ if it is compact and otherwise arbitrary positive. $M$ reduces to the ADM (Arnowitt-Deser-Misner) mass at spacelike infinity in the asymptotically flat spacetime. In terms of $M$, some components of the Lovelock equation are written in the following simple form~\cite{mwr2011,tw2011}: \begin{align} D_A M =&V_{n-2}^{(k)}R^{n-2}\biggl({T_A}^B(D_B R) -{T^B}_B (D_A R)\biggl). \label{1stlaw1} \end{align} \subsection{Vacuum solutions} In the vacuum case ($T_{\mu\nu}=0$), Eq.~(\ref{1stlaw1}) shows that $M$ is constant. The maximally symmetric solution, namely Minkowski, de~Sitter (dS) or anti-de~Sitter (AdS) solution, gives $M=0$. The maximally symmetric spacetime may be given in the following coordinates: \begin{align} ds^2=&-(k-{\tilde \lambda}r^2)dt^2+\frac{dr^2}{k-{\tilde \lambda}r^2}+r^2\gamma_{ab}dz^a dz^b,\label{vacuum} \end{align} where ${\tilde \lambda}:=2\lambda/[(n-1)(n-2)]$ and $\lambda$ is the effective cosmological constant, which is determined by the following algebraic equation: \begin{align} \label{lambda} 0=\sum_{p=0}^{[n/2]}{\tilde \alpha}_{(p)}{\tilde \lambda}^p=:v({\tilde\lambda}). \end{align} The Minkowski vacuum ($\lambda=0$) is possible only if ${\alpha}_{(0)}=0$. Since Eq.~(\ref{lambda}) is a higher-order polynomial, there can be multiple values of ${\tilde \lambda}$. We call the vacuum ${\tilde \lambda}={\tilde \lambda}_1$ {\it non-degenerate} if $(dv/d{\tilde\lambda})({\tilde \lambda}_1)\ne 0$ holds. A {\it simply} degenerate vacuum is characterized by $(dv/d{\tilde\lambda})({\tilde \lambda}_1)= 0$, while a {\it doubly} degenerate vacuum is characterized by $(dv/d{\tilde\lambda})({\tilde \lambda}_1)= (d^2v/d{\tilde\lambda}^2)({\tilde \lambda}_1)= 0$. In a similar manner, a $q$th-order degenerate vacuum is defined by $(d^sv/d{\tilde\lambda}^s)({\tilde \lambda}_1)=0$ for $s=1,2,\cdots,q$, where $q \le [(n-3)/2]$ is satisfied because of ${\tilde \alpha}_{(n/2)}\equiv 0$ for even $n$. The Schwarzschild-Tangherlini-type vacuum solution in Lovelock gravity~\cite{zegers2005} is given by \begin{align} ds^2=-f(r)dt^2+\frac{dr^2}{f(r)}+r^2\gamma_{ab}dz^a dz^b, \label{f-vacuum} \end{align} where the metric function $f(r)$ is determined algebraically by \begin{align} \label{alg} {\tilde M} =\sum_{p=0}^{[n/2]}{\tilde \alpha}_{(p)}r^{n-1-2p}(k-f(r))^p. \end{align} ${\tilde M}$ is related to the constant generalized Misner-Sharp mass as ${\tilde M}:=2\kappa_n^2M/[(n-2)V_{n-2}^{(k)}]$. This class of vacuum solutions in higher-order Lovelock gravity was first obtained by Boulware and Deser~\cite{bdw} in the quadratic theory. (See~\cite{bdw2} for further discussions.) The above solution reduces to the ones found in~\cite{DCBH,DCBH2} in the case where the coupling constants are chosen such that the theory admits a fully degenerate maximally symmetric vacuum. Birkhoff's theorem in Lovelock gravity asserts that, in the case where $(DR)^2:=(D_AR)(D^AR)\ne 0$ and the spacetime is of the $C^2$-class, there is a unique vacuum solution as long as the theory does not admit degenerate vacua~\cite{zegers2005}. (See also~\cite{mwr2011} for more general case.) One of the purposes of the present paper is to derive the formulae to quantize a Lovelock black hole described by the above solution. If the theory admits degenerate vacua, there may be more vacuum solutions. If the theory admits simply degenerate vacua, the following is also a vacuum solution: \begin{align} ds^2=&-(k-{\tilde\lambda} r^2)e^{2\delta(t,r)}dt^2+\frac{dr^2}{k-{\tilde\lambda} r^2}+r^2\gamma_{ab}dz^a dz^b, \label{type-I} \end{align} where $\delta(t,r)$ is an {\it arbitrary} function and ${\tilde\lambda}$ takes the value for the degenerate vacuum. (This solution was first obtained properly in quadratic Lovelock gravity by Charmousis and Dufaux~\cite{cd2002}.) If the theory admits doubly degenerate vacua, there is another vacuum solution where the two-dimensional portion $(M^2, g_{AB})$ is {\it totally arbitrary}~\cite{mwr2011}. {If one removes the assumption of $C^2$-differentiability of the spacetime, then more vacuum solutions exist~\cite{Garraffo:2007fi,Gravanis:2010zs}.} \section{Dimensionally reduced action} \label{Dim red action} \subsection{Covariant form} In the symmetric spacetime under consideration, the system may be described by the effective two-dimensional action: \begin{align} I_{(2)}=I_{M}+I_{\partial {M}}.\label{action3} \end{align} The dynamical term $I_{M}$ is written as \begin{align} I_{M}=\int d{\bar y}^0L[g_{AB},R]=\int d{\bar y}^0 \int d{\bar y}^1{\cal L}[g_{AB},R],\label{2-action} \end{align} where ${\bar y}^0$ is a timelike coordinate on $(M^2, g_{AB})$. Here the Lagrangian $L$ and the Lagrangian density ${\cal L}$ are functionals of the metric functions, which are determined up to a total derivative. The main purpose of geometrodynamics is to find canonical variables (that are functionals of the metric functions) to provide a tractable form and transparent physical meaning for ${\cal L}$. Our first task is to derive a tractable tensorial form of $I_{M}$. For symmetric spacetimes under consideration, the action reduces to \begin{align} I_{M}=\frac{V_{n-2}^{(k)}}{2\kappa_n^2}\int d^2{\bar y}\sqrt{-g_{(2)}}R^{n-2}\sum^{[n/2]}_{p=0}\alpha_{(p)} {\cal L}_{(p)}, \label{eq:reduced action 2} \end{align} where $g_{(2)}:=\det(g_{AB})$ and the dimensionally reduced $p$th-order Lovelock term $ {\cal L}_{(p)}$ is given from expressions (2.19) and (2.20) of~\cite{mwr2011} as \begin{align} {\cal L}_{(p)} =& \frac{(n-2)!}{(n-2p)!} \biggl[(n-2p)(n-2p-1)\left(\frac{k-(DR)^2}{R^2}\right)^{p} - 2p(n-2p)\frac{D^2R}{R}\left(\frac{k-(DR)^2}{R^2}\right)^{p-1} \nonumber \\ & + 2p(p-1) \frac{(D^2R)^2 - (D^AD_BR)(D^BD_AR)}{R^2} \left(\frac{k-(DR)^2}{R^2}\right)^{p-2} + p\overset{(2)}{\cal R} \left(\frac{k-(DR)^2}{R^2}\right)^{p-1} \biggl], \label{L_p} \end{align} where ${}^{(2)}{\cal R}$ is the Ricci scalar on $(M^2, g_{AB})$ and $D^2R:=D^AD_AR$. At a glance, there is a non-minimal coupling between $(DR)^2$ and ${}^{(2)}{\cal R}$ in ${\cal L}_{(p)}$. Such a Lagrangian is not tractable to perform the canonical analysis. As proven in Appendix~\ref{appendix1}, we can write it, up to total divergences, without such a coupling: \begin{align} \label{eq:lovelock simplified} {\cal L}_{(p)} =& \frac{(n-2)!}{(n-2p)!} \Biggl[pk^{p-1}\overset{(2)}{\cal R} R^{2-2p} + pR^{2-n}\frac{D^A(R^{n-2p})D_A((DR)^2)}{(DR)^2}\biggl\{k^{p-1}-(k-(DR)^2)^{p-1}\biggl\} \nonumber \\ & + (n-2p)(n-2p-1)\biggl\{\left(k-(DR)^2\right)^{p} +2pk^{p-1}(DR)^2\biggl\}R^{-2p} \Biggr]. \end{align} This is a key result of our paper and the starting point of our canonical analysis. Note that in the following we work exclusively with the equations of motion derived from the reduced action (\ref{eq:reduced action 2}). In general it is not true that dimensional reduction commutes with the variational principle. That is, the space of extrema of a dimensionally reduced action in principle may not coincide with the space of symmetric solutions of the unreduced action. However, in a very elegant and powerful set of papers \cite{Palais1979,Fels2002} (see also \cite{Deser2003}), it has been rigorously proven that if the symmetry group is a compact Lie group, as in our case, then for any local metric theory of gravity in arbitrary space-time dimensions, with or without matter, variation does indeed commute with dimensional reduction. The spherically symmetric equations of motion obtained from the full, unreduced Lovelock action with matter were explicitly written down in \cite{mwr2011}. The proof that the solution space is the same in both cases nonetheless requires the more detailed analysis of \cite{Palais1979,Fels2002}. Unfortunately, this analysis is only valid in the compact case, so that more work needs to be done in order to prove that the dimensionally reduced action is sufficient when $k=0,-1$. This is one of the reasons that we defer consideration of the non-compact case to a future study. \subsection{ADM form} We are going to write down the action (\ref{eq:lovelock simplified}) by adopting the following ADM coordinates $(t,x)$ on $(M^2,g_{AB})$: \begin{eqnarray} ds_{(2)}^2=g_{AB}d{\bar y}^Ad{\bar y}^B=-N(t,x)^2dt^2+\Lambda(t,x)^2(dx+N_r(t,x)dt)^2.\label{ADM} \end{eqnarray} Now canonical variables are $N$, $N_r$, $\Lambda$, and $R$ and their momentum conjugates are respectively written as $P_{N}$, $P_{N_r}$, $P_{\Lambda}$, and $P_{R}$. In the present paper, a dot and a prime denote a partial derivative with respect to $t$ and $x$, respectively. The metric and its inverse are \begin{align} g_{tt}=&-(N^2-\Lambda^2N_r^2),\quad g_{tx}=\Lambda^2N_r,\quad g_{xx}=\Lambda^2,\\ g^{tt}=&-N^{-2},\quad g^{tx}=N_rN^{-2},\quad g^{xx}=N^{-2}\Lambda^{-2}(N^2-\Lambda^2N_r^2), \end{align} while $\sqrt{-g_{(2)}}$ is given by \begin{align} \sqrt{-g_{(2)}}=N\Lambda. \end{align} For the later use, we compute the following quantities: \begin{align} F:=&(DR)^2 \nonumber \\ =&-y^2+\Lambda^{-2}{R'}^2,\label{defF}\\ \sqrt{-g_{(2)}}D^2 R=&-\partial_t(\Lambda y)+\partial_x(\Lambda N_r y+\Lambda^{-1}NR'), \end{align} where $y$ is defined by \be y:=N^{-1}({\dot R}-N_rR').\label{defy} \ee We also need the following relationship: \bea D^A(R^{n-2p})D_A((DR)^2) &=& (n-2p) R^{n-2p-1} \left( - \frac{1}{N}y \dot{F} + \left(\frac{R'}{\Lambda^2} +\frac{N_r}{N} y\right) F'\right). \eea Using this result, the action (\ref{eq:lovelock simplified}) is written in the following simple form: \begin{align} I_M=&\frac{(n-2)V_{n-2}^{(k)}}{2\kappa_n^2}\sum^{[n/2]}_{p=0}\int d^2{\bar y}\sqrt{-g_{(2)}}\frac{{\tilde \alpha}_{(p)} }{(n-2p)} \Biggl[pk^{p-1}\overset{(2)}{\cal R}R^{n-2p} \nonumber \\ &- p(n-2p)\frac{R^{n-2p-1}}{N\Lambda}\{k^{p-1}-(k-F)^{p-1}\}\biggl\{\Lambda y\frac{\dot F}{F}-(\Lambda N_r y+\Lambda^{-1}NR')\frac{F'}{F}\biggl\} \nonumber \\ & + (n-2p)(n-2p-1)\biggl\{\left(k-F\right)^{p} +2pk^{p-1}F\biggl\}R^{n-2-2p} \Biggr]. \label{action-3} \end{align} The first term is the two-dimensional gravity non-minimally coupled scalar field $R$, which is essentially the same as the general relativistic case. This term can be explicitly written down in terms of the canonical variables using \begin{align} \sqrt{-g_{(2)}}R^{n-2p}\overset{(2)}{\cal R}=&-2N^{-1}\biggl((R^{n-2p})'N_r-\partial_t(R^{n-2p})\biggl)(N_r'\Lambda +N_r\Lambda'-{\dot \Lambda}) \nonumber \\ &-2N\biggl((R^{n-2p})''\Lambda^{-1}+(R^{n-2p})'(\Lambda^{-1})'\biggl)+\partial_t(\cdots)+\partial_x (\cdots).\label{intR} \end{align} Based on the action (\ref{action-3}), we will perform the canonical analysis in the subsequent sections using geometrodynamical phase space variables. We therefore now review briefly the geometrodynamics of Kucha\v{r}~\cite{kuchar94}. \subsection{Geometrodynamics} The metric (\ref{ADM}) may be written in the generalized Schwarzschild form in terms of the areal coordinates as \begin{align} ds_{(2)}^2=-F(R,T)e^{2\sigma(R,T)}dT^2+F(R,T)^{-1}dR^2. \label{eq:Schwarzschild metric} \end{align} The generalized Misner-Sharp mass $M$ is then given by \begin{align} M(R,T)= \frac{(n-2)V_{n-2}^{(k)}}{2\kappa_n^2}\sum_{p=0}^{[n/2]}{\tilde \alpha}_{(p)}R^{n-1-2p}\biggl(k-F(R,T)\biggl)^p.\label{M-F} \end{align} This implicitly gives the functional form $F=F(R,M)$. However, there is no one-to-one correspondence between $F$ and $M$ unless all the coupling constants $\alpha_{(p)}$ are non-negative. To see the relation to the ADM form (\ref{ADM}) we use the coordinate transformations $T=T(t,x)$ and $R=R(t,x)$, to write the metric (\ref{eq:Schwarzschild metric}) as \begin{align} ds_{(2)}^2=&-(F{\dot T}^2e^{2\sigma}-F^{-1}{\dot R}^2)dt^2+2(-F{\dot T}T'e^{2\sigma}+F^{-1}{\dot R}R')dtdx \nonumber \\ &+(-F{T'}^2e^{2\sigma}+F^{-1}{R'}^2)dx^2. \end{align} Comparing with the ADM form, we identify \begin{align} F{\dot T}^2e^{2\sigma}-F^{-1}{\dot R}^2=&N^2-\Lambda^2{N_r}^2,\\ -F{\dot T}T'e^{2\sigma}+F^{-1}{\dot R}R'=&\Lambda^2N_r,\\ -F{T'}^2e^{2\sigma}+F^{-1}{R'}^2=&\Lambda^2 \label{Gamma1} \end{align} and obtain \begin{align} N_r=&\frac{-F{\dot T}T'e^{2\sigma}+F^{-1}{\dot R}R'}{-F{T'}^2e^{2\sigma}+F^{-1}{R'}^2},\label{beta}\\ N=& \frac{e^{\sigma}({\dot T}R'-{\dot R}T')}{\sqrt{-F{T'}^2e^{2\sigma}+F^{-1}{R'}^2}},\label{alpha} \end{align} { As discussed by Kucha\v{r} in Section IVA of~\cite{kuchar94}, one can ensure that ${\dot T}R'-{\dot R}T'$ and hence the Lapse function $N$ are positive by an appropriate choice of $x$.} $y$ is then given from the definition (\ref{defy}) as \begin{align} y=\frac{FT'e^{\sigma}}{\sqrt{-F{T'}^2e^{2\sigma}+F^{-1}{R'}^2}}, \end{align} from which we obtain \begin{align} T'e^{\sigma}=\frac{y\Lambda}{F},\label{T-sigma} \end{align} where we used Eq.~(\ref{Gamma1}). Using this to eliminate $T'e^{\sigma}$ in Eq.~(\ref{Gamma1}), we obtain \begin{align} F=-y^2+\frac{{R'}^2}{\Lambda^{2}} \label{eq:F1} \end{align} as required by consistency with (\ref{defF}) In the above, we derived expressions for the generalized Schwarzschild time $T$ in terms of the canonical ADM variables. As we will see in the following this determines the conjugate momentum to the Misner-Sharp mass function in a form that is appropriate for slicings that approach the Schwarzschild form at spatial infinity. Other asymptotic forms for the slicings are possible, including flat slice or generalized Painlev\'{e}-Gullstrand (PG) coordinates: \be ds_{(2)}^2=-{e}^{2\sigma}dT_{\rm PG}^2 + (dR + G{e}^{\sigma} dT_{\rm PG})^2, \label{p-g-metric} \ee where $\sigma=\sigma(T_{\rm PG},R)$ and $G=G(T_{\rm PG},R)$. The geometrodynamical variables appropriate for such slicings were first derived in \cite{Louko2007}. Since we have \be (D R)^2=1-G^2 \ee for the above form of the metric, it follows that \be G=\pm\sqrt{1-F}. \ee By inspection of (\ref{p-g-metric}) one can see that the positive sign yields an equation for ingoing null geodesics that is regular at any horizon $F=0$, so this is the choice that is suitable for describing the spacetime near a future horizon (black hole). The opposite sign must be chosen for a past horizon (white hole). We now go through exactly the same derivation as before. {Performing the coordinate transformations $T_{\rm PG}=T_{\rm PG}(t,x)$ and $R=R(t,x)$ in the metric (\ref{p-g-metric}) and comparing to the ADM form (\ref{ADM}) yields: } \begin{subequations \label{eq:rec1 \bea \Lambda^2 &=& (R'+{e}^{\sigma} G T_{\rm PG}')^2 -{e}^{2\sigma}{T_{\rm PG}'}^2, \label{conformal-factor} \\ N^2-\Lambda^2N_r^2 &=&{e}^{2\sigma} \dot{T}_{\rm PG}^2 - (\dot{R}+{e}^{\sigma} G \dot{T}_{\rm PG})^2, \label{eq:mixed} \\ \Lambda^2 N_r &=& (R'+{e}^{\sigma} G T_{\rm PG}')(\dot{R}+{e}^{\sigma} G \dot{T}_{\rm PG}) -{e}^{2\sigma}T_{\rm PG}'\dot{T}_{\rm PG}. \label{shift} \eea \end{subequations} Solving (\ref{eq:rec1}) for $N$ and~$N_r$, we fin \begin{subequations \label{eq:N-and-sigma-solved \bea N_r &=& \frac{(R'+{e}^{\sigma} G T_{\rm PG}')(\dot{R}+{e}^{\sigma} G \dot{T}_{\rm PG}) -{e}^{2\sigma}T_{\rm PG}'\dot{T}_{\rm PG}}{(R'+{e}^{\sigma} G T_{\rm PG}')^2-{{e}^{\sigma}(T_{\rm PG}')}^2} , \\ N &=& \frac{R'{e}^{\sigma}\dot{T}_{\rm PG}- \dot{R}{e}^{\sigma} T_{\rm PG}'}{\sqrt{(R'+{e}^{\sigma} G T_{\rm PG}')^2-{{e}^{\sigma}(T_{\rm PG}')}^2}} . \label{eq:sigma-solved \eea \end{subequations} To complete the derivation, we use (\ref{defy}) and the above expressions for $\Lambda$, $N$ and $N_r$ to calculate \bea y\Lambda &=& (1-G^2)e^{\sigma}T_{\rm PG}' + G R', \label{eq:yLambda PG} \eea which yields: \be e^{\sigma}T_{\rm PG}' = \frac{y \Lambda}{F} \pm \frac{R'\sqrt{1-F}}{F}. \label{eq:Tprime PG} \ee The second term on the right-hand side of the above guarantees that the PG time is well defined either for (with a +ve sign) future or (with a -ve sign) past horizons. \section{Canonical formalism in general relativity} \label{Can formalism gr} In this section, we perform the canonical analysis for spherically symmetric spacetimes ($k=1$) in general relativity without a cosmological constant, which is a generalization of the Kucha\v{r}'s analysis in four dimensions to arbitrary dimensions. We set ${\tilde\alpha}_{(1)}=1$ in this section for simplicity. The reduced action (\ref{action-3}) then becomes quite simple: \begin{align} I_{\rm M(GR)}=&\frac{\ma A_{n-2}}{2\kappa_n^2}\int d^2{\bar y} \Biggl[2(n-2)R^{n-3}y(N_r'\Lambda +N_r\Lambda') -2N\biggl((R^{n-2})''\Lambda^{-1}+(R^{n-2})'(\Lambda^{-1})'\biggl) \nonumber \\ & + (n-2)(n-3)(1+F)N\Lambda R^{n-4}-2(n-2)R^{n-3}y {\dot \Lambda}\Biggr]. \label{action-3gr} \end{align} Here $\ma A_{n-2}$ is the surface area of an $(n-2)$-dimensional unit sphere, namely \begin{align} \ma A_{n-2} :=\frac{2\pi^{(n-1)/2}}{\Gamma((n-1)/2)}(\equiv V_{n-2}^{(1)}), \label{unitarea} \end{align} where $\Gamma(x)$ is the Gamma function. The purpose of this section is to show that the areal radius and the Misner-Sharp mass are well-defined canonical variables in the system, which will be generalized to Lovelock gravity in the following . \subsection{ADM variables} We first derive the expressions for $P_\Lambda$ and $P_R$. The corresponding Lagrangian density of the action (\ref{action-3gr}) is \begin{align} {\cal L}=& \frac{(n-2)\ma A_{n-2}}{2\kappa_n^2} \Biggl[2R^{n-3}y(N_r'\Lambda +N_r\Lambda')-\frac{2}{(n-2)}N\biggl((R^{n-2})''\Lambda^{-1}+(R^{n-2})'(\Lambda^{-1})'\biggl) \nonumber \\ &+ (n-3)(1+F)N\Lambda R^{n-4}-2R^{n-3}y {\dot \Lambda}\biggl], \end{align} from which we obtain $P_N=P_{N_r}=0$ and \begin{align} P_\Lambda=& -\frac{(n-2)\ma A_{n-2}}{\kappa_n^2}R^{n-3}N^{-1}({\dot R}-N_rR'),\label{PLambdaGR2}\\ P_R=& \frac{(n-2)\ma A_{n-2} }{2\kappa_n^2} \Biggl[2R^{n-3}N^{-1}(N_r'\Lambda +N_r\Lambda')-2 (n-3)N^{-1}\Lambda R^{n-4}({\dot R}-N_rR')-2R^{n-3}N^{-1} {\dot \Lambda}\biggl].\label{PR} \end{align} With $P_\Lambda$ and $P_R$, the Hamiltonian density ${\cal H}(:={\dot \Lambda}P_{\Lambda}+{\dot R}P_{R}-{\cal L})$ is given by \begin{align} {\cal H}=&(N_r\Lambda P_{\Lambda})'-N_r\Lambda P_{\Lambda}'+N_rR'P_{R}-\frac{\kappa_n^2 N}{(n-2)\ma A_{n-2}R^{n-2}}P_{\Lambda}\biggl(RP_{R}-\frac{n-3}{2}\Lambda P_{\Lambda}\biggl) \nonumber \\ &-\frac{(n-2)\ma A_{n-2}}{\kappa_n^2}N\biggl\{R^{n-3}\biggl(-R'' \Lambda^{-1}+R'\Lambda^{-2}\Lambda'\biggl)+\frac{n-3}{2} \Lambda R^{n-4}\biggl(1-\Lambda^{-2}{R'}^2\biggl)\biggl\}. \label{hamildens-gr} \end{align} The first term in Eq.~(\ref{hamildens-gr}) is the total derivative and becomes a boundary term. Since $P_N=P_{N_r}=0$, the Hamilton equations for $N$ and $N_r$ give constraint equations $H=0$ and $H_r=0$, where the super-momentum $H_{r}(:=\delta {\cal H}/\delta N_r)$ and the super-Hamiltonian $H(:=\delta {\cal H}/\delta N)$ are given by \begin{align} H_r=&-\Lambda P_{\Lambda}'+R'P_{R}, \label{Hr}\\ H=&-\frac{\kappa_n^2}{(n-2)\ma A_{n-2}R^{n-2}}P_{\Lambda}\biggl(RP_{R}-\frac{n-3}{2}\Lambda P_{\Lambda}\biggl) \nonumber \\ &-\frac{(n-2)\ma A_{n-2}}{\kappa_n^2}\biggl\{R^{n-3}\biggl(-R'' \Lambda^{-1}+R'\Lambda^{-2}\Lambda'\biggl)+\frac{n-3}{2} \Lambda R^{n-4}\biggl(1-\Lambda^{-2}{R'}^2\biggl)\biggl\}. \label{H} \end{align} The action is finally written as \begin{align} I_{\rm M(GR)}=&\int dt\int dx ( {\dot \Lambda}P_{\Lambda}+{\dot R}P_{R}-N H-N_r H_{r}).\label{IMGR} \end{align} It can be verified that with suitable boundary conditions the constraints $H$ and $H_r$ are first class in the Dirac sense and generate spacetime diffeomorphisms that preserve the spherically symmetric form of the metric. \subsection{The Schwarzschild-Tangherlini spacetime in various coordinate systems} In this subsection, we review various coordinate systems in the Schwarzschild-Tangherlini spacetime. The Schwarzschild-Tangherlini vacuum solution in the best-known Schwarzschild coordinates is given by \begin{align} ds^2=&-\biggl(1-\frac{{\tilde M}}{r^{n-3}}\biggl)dt^2+\biggl(1-\frac{{\tilde M}}{r^{n-3}}\biggl)^{-1}dr^2+r^2\gamma_{ab}dz^a dz^b, \label{st} \end{align} where ${\tilde M}:=2\kappa_n^2M/[(n-2){\cal A}_{n-2}]$ and $M$ is the ADM mass. The corresponding ADM variables are \begin{align} N^2=1-\frac{{\tilde M}}{r^{n-3}},\quad N_r=0,\quad \Lambda^2=&\biggl(1-\frac{{\tilde M}}{r^{n-3}}\biggl)^{-1}, \quad R=r. \end{align} In the next subsection, we will consider the boundary condition at spacelike infinity with this slicing. However, there is a variety of slicings in the Schwarzschild-Tangherlini spacetime, as presented below. By introducing a new spacelike coordinate $\rho$ as $r=\rho[1+{\tilde M}/(4\rho^{n-3})]^{2/(n-3)}$, the metric~(\ref{st}) is transformed into the isotropic coordinates: \begin{align} ds^2=&-\frac{[1-{\tilde M}/(4\rho^{n-3})]^2}{[1+{\tilde M}/(4\rho^{n-3})]^2}dt^2+\left(1+\frac{{\tilde M}}{4\rho^{n-3}}\right)^{4/(n-3)}(d\rho^2+\rho^2\gamma_{ab}dz^a dz^b). \label{Sch-iso} \end{align} The corresponding ADM variables are \begin{align} N^2=&\frac{[1-{\tilde M}/(4\rho^{n-3})]^2}{[1+{\tilde M}/(4\rho^{n-3})]^2},\quad N_r=0,\quad \Lambda^2=\left(1+\frac{{\tilde M}}{4\rho^{n-3}}\right)^{4/(n-3)},\quad R=\rho\left(1+\frac{{\tilde M}}{4\rho^{n-3}}\right)^{2/(n-3)}. \end{align} On the other hand, by introducing a new time coordinate $\tau$ defined by \begin{equation} d\tau:=dt+\sqrt{\frac{{\tilde M}}{r^{n-3}}}\frac{dr}{(1-{\tilde M}/r^{n-3})}, \end{equation} the metric~(\ref{st}) is transformed into the Painlev\'{e}-Gullstrand coordinates: \begin{equation} ds^2=-\biggl(1-\frac{{\tilde M}}{r^{n-3}}\biggl)d\tau^2+2\sqrt{\frac{{\tilde M}}{r^{n-3}}}d\tau dr+dr^2+r^2\gamma_{ab}dz^a dz^b. \label{Sch-P} \end{equation} The corresponding ADM variables are \begin{align} N^2=1,\quad N_r=\sqrt{\frac{{\tilde M}}{r^{n-3}}},\quad \Lambda^2=1,\quad R=r. \end{align} By the coordinate transformation \begin{equation} r=\biggl(\frac{n-1}{2}\biggl)^{4/(n-1)}{\tilde M}^{1/(n-1)}\biggl(\frac{2({\tilde\rho}-\tau)}{n-1}\biggl)^{2/(n-1)} \end{equation} from the Painlev\'{e}-Gullstrand coordinates (\ref{Sch-P}), we obtain the Schwarzschild-Tangherlini metric in the Lema\^{\i}tre coordinates: \begin{align} ds^2=&-d\tau^2+{\tilde M}^{2/(n-1)}\biggl[\biggl(\frac{n-1}{2}({\tilde\rho}-\tau)\biggl)^{-2(n-3)/(n-1)}d{\tilde\rho}^2+\biggl(\frac{n-1}{2}({\tilde\rho}-\tau)\biggl)^{4/(n-1)}\gamma_{ab}dz^a dz^b\biggl]. \label{Sch-L} \end{align} In this coordinate system, the central curvature singularity and the black-hole event horizon are represented by $\tau={\tilde\rho}$ and \begin{align} {\tilde\rho}-\tau=\frac{2{\tilde M}^{1/(n-3)}}{n-1}, \end{align} respectively. In the Lema\^{\i}tre coordinates, the radial coordinate ${\tilde\rho}$ does not coincide with $R$ in the asymptotically flat region. Defining a new radial coordinate as \begin{align} {\tilde r}:={\tilde M}^{1/(n-1)}\biggl(\frac{n-1}{2}{\tilde\rho}\biggl)^{2/(n-1)}, \end{align} which coincides with $R$ in the asymptotic region, we transform the metric (\ref{Sch-L}) into the following form: \begin{align} ds^2=&-d\tau^2+\biggl(1-\frac{(n-1){\tilde M}^{1/2}\tau}{2{\tilde r}^{(n-1)/2}}\biggl)^{-2(n-3)/(n-1)}d{\tilde r}^2 \nonumber\\ & +{\tilde r}^2\biggl(1-\frac{(n-1){\tilde M}^{1/2}\tau}{2{\tilde r}^{(n-1)/2}}\biggl)^{4/(n-1)}\gamma_{ab}dz^a dz^b. \label{Sch-L2} \end{align} This metric is inhomogeneous and time-dependent in this coordinate system and the corresponding ADM variables and their fall-off rates are \begin{align} N^2=&1,\quad N_r=0,\nonumber\\ \Lambda^2=&\biggl(1-\frac{(n-1){\tilde M}^{1/2}\tau}{2{\tilde r}^{(n-1)/2}}\biggl)^{-2(n-3)/(n-1)},\quad R={\tilde r}\biggl(1-\frac{(n-1){\tilde M}^{1/2}\tau}{2{\tilde r}^{(n-1)/2}}\biggl)^{2/(n-1)}. \end{align} Lastly, we present the Schwarzschild-Tangherlini metric in the Kerr-Schild coordinates: \begin{align} ds^2=&-d{\hat t}^2+dr^2+r^2\gamma_{ab}dz^a dz^b+\frac{{\tilde M}}{r^{n-3}}(d{\hat t}+dr)^2 \label{Sch-SD}\\ =&-\biggl(1-\frac{{\tilde M}}{r^{n-3}}\biggl)d{\hat t}^2+\frac{2{\tilde M}}{r^{n-3}}d{\hat t}dr+\biggl(1+\frac{{\tilde M}}{r^{n-3}}\biggl)dr^2+r^2\gamma_{ab}dz^a dz^b. \label{Sc-6} \end{align} The corresponding ADM variables are \begin{align} N^2=&\biggl(1+\frac{{\tilde M}}{r^{n-3}}\biggl)^{-1},\quad N_r=\frac{{\tilde M}}{r^{n-3}}\biggl(1+\frac{{\tilde M}}{r^{n-3}}\biggl)^{-1},\quad \Lambda^2=1+\frac{{\tilde M}}{r^{n-3}},\quad R=r. \end{align} \subsection{Boundary condition and boundary terms} To perform the geometrodynamics, the boundary condition plays a crucial role. In the present paper, we adopt the following boundary condition at spacelike infinity $x\to \pm \infty$\footnote{These boundary conditions are suited to asymptotically Schwarzschild slicings. The analogous boundary conditions for PG coordinates are given in \cite{Husain} and discussed in Appendix~\ref{app:boundary}.}: \begin{align} N\simeq& N_\infty(t)+\mathcal{O}(x^{-\epsilon_1}),\label{bc1}\\ N_r\simeq& N_r^\infty(t) x^{-(n-3)/2-\epsilon_2},\label{bc2}\\ \Lambda \simeq& 1+\Lambda_1(t)x^{-(n-3)},\label{bc3}\\ R\simeq& x+R_1(t)x^{-(n-4)-\epsilon_4},\label{bc4} \end{align} where $\epsilon_1$ is a positive number and $\epsilon_2$ and $\epsilon_4$ satisfy $\epsilon_2> \max[0,-(n-5)/2]$ and $\epsilon_4>\max[0,-(n-5)]$. { (The validity of this boundary condition is verified in Appendix~\ref{app:boundary}.)} The asymptotic behavior of $P_\Lambda$ and $P_R$ are given by \begin{align} P_\Lambda\simeq& -\frac{(n-2)\ma A_{n-2}}{\kappa_n^2}N_\infty^{-1}\biggl({\dot R_1} x^{1-\epsilon_4}-N_r^{\infty} x^{(n-3)/2-\epsilon_2}\biggl),\label{bc5}\\ P_R\simeq & -\frac{(n-2)\ma A_{n-2} }{\kappa_n^2} N_\infty^{-1}\biggl[N_r^{\infty}(t) \biggl(-\frac{n-3}{2}+\epsilon_2\biggl)x^{(n-5)/2-\epsilon_2}+{\dot \Lambda}_1(t)\biggl].\label{bc6} \end{align} Under the boundary condition adopted, the Misner-Sharp mass converges to a finite value $M\simeq M^\infty(t)$, where $M^\infty(t)$ is related to $\Lambda_1(t)$ as \begin{align} \Lambda_1(t)\equiv \frac{\kappa_n^2M^\infty(t)}{(n-2)\ma A_{n-2}}.\label{bc7} \end{align} Now let us consider the boundary term for the action (\ref{IMGR}). The role of the boundary term is to subtract the diverging terms at the boundary in the variation of the above action. The action is completed by adding the boundary term, which gives a finite value in the variation. Since the variation of $I_{\rm M(GR)}$ gives \begin{align} \delta I_{\rm M(GR)}=&\int dt\int dx \biggl( \partial_t (\delta {\Lambda}P_{\Lambda})-\delta \Lambda{\dot P_{\Lambda}}+ {\dot \Lambda}\delta P_{\Lambda}+ \partial_t (\delta {R}P_{R})-\delta R{\dot P_{R}}+{\dot R}\delta P_{R} \nonumber \\ &-\delta N H-N \delta H-\delta N_r H_{r}-N_r \delta H_{r}\biggl),\label{variation1gr} \end{align} we need to know the contributions from $N_r\delta H_r$ and $N\delta H$. Using the following results; \begin{align} N_r\delta H_r=&-N_r\delta \Lambda P_{\Lambda}'-(N_r\Lambda\delta P_\Lambda)'+(N_r\Lambda)' \delta P_{\Lambda}+(N_r\delta RP_R)'-\delta R(N_rP_{R})'+N_rR'\delta P_{R},\\ N\delta H=&\biggl(\mbox{irrelevant terms}\biggl)-\frac{(n-2)\ma A_{n-2}}{\kappa_n^2}\biggl\{-\biggl((NR^{n-3}\Lambda^{-1}\delta R)'-(NR^{n-3}\Lambda^{-1})'\delta R\biggl)' \nonumber \\ &+(N'R^{n-3}\Lambda^{-1}\delta R)'-(N'R^{n-3}\Lambda^{-1})'\delta R +(NR^{n-3}R'\Lambda^{-2}\delta \Lambda)'-(NR^{n-3}R'\Lambda^{-2})'\delta \Lambda\biggl\}, \end{align} we can write (\ref{variation1gr}) in the following form: \begin{align} \delta I_{\rm M(GR)}=&\int dt\int dx \biggl(\mbox{dynamical terms}\biggl)+\int dx \biggl[\delta {\Lambda}P_{\Lambda}+\delta {R}P_{R}\biggl]_{t=t_1}^{t=t_2} \nonumber \\ &-\int dt\biggl[-N_r\Lambda\delta P_\Lambda+N_rP_R\delta R \nonumber \\ &-\frac{(n-2)\ma A_{n-2}}{\kappa_n^2}\biggl\{-NR^{n-3}\Lambda^{-1}\delta (R')+N'R^{n-3}\Lambda^{-1}\delta R+NR^{n-3}R'\Lambda^{-2}\delta \Lambda\biggl\}\biggl]_{x=-\infty}^{x=+\infty}. \end{align} Now the boundary condition comes into play. We assume $\delta \Lambda=\delta R=0$ at $t=t_1,t_2$ and then the second term in the above variation vanishes. Using the boundary condition (\ref{bc1})--(\ref{bc6}), we can show that only the contribution in the last integral comes from $NR^{n-3}R'\Lambda^{-2}\delta \Lambda$ as \begin{align} NR^{n-3}R'\Lambda^{-2}\delta \Lambda\simeq N_\infty\delta \Lambda_1= \frac{\kappa_n^2N_\infty\delta M^\infty(t)}{(n-2)\ma A_{n-2}}, \end{align} where we used Eq.~(\ref{bc7}). Finally we obtain the boundary term in a simple form: \begin{align} \delta I_{\rm M(GR)}=&\int dt\int dx \biggl(\mbox{dynamical terms}\biggl)+\int dt\biggl[N_\infty(t)\delta M^\infty(t)\biggl]_{x=-\infty}^{x=+\infty}. \label{boundaryGR} \end{align} \subsection{Misner-Sharp mass as canonical variable} In the ADM coordinates, the canonical variables are $\{\Lambda, P_\Lambda; R,P_R\}$. However, the physical meanings of the variable $\Lambda$ is not so clear. In this subsection, we show that the two-dimensional equivalent action is written in a rather elegant manner by introducing the Misner-Sharp mass $M$ as a canonical variable. We introduce a new set of canonical variables $\{M, P_M; S,P_S\}$ defined by \begin{align} S:=&R, \label{eq:R}\\ P_S:=&P_R-\frac{1}{R'}(\Lambda P_\Lambda'+P_MM'), \label{eq:PM}\\ M :=& \frac{(n-2)\ma A_{n-2}}{2\kappa_n^2}R^{n-3}(1-F),\label{MS}\\ P_M:=&-T'e^{\sigma}=-\frac{y\Lambda}{F}, \label{eq:Pr1} \end{align} where we used Eq.~(\ref{T-sigma}). We are going to show below that, under the boundary condition (\ref{bc1})--(\ref{bc6}), the transformation from a set of variables $\{\Lambda, P_\Lambda; R, P_R\}$ to another set $\{M, P_M; S, P_S\}$ is a well-defined canonical transformation. Note (\ref{eq:Pr1}) chooses the conjugate to $M$ in terms of the Schwarzschild time $T$. As verified in Appendix~\ref{app:boundary} this leads to a finite Liouville form providing one chooses boundary conditions such that the metric approaches the vacuum Schwarzschild solution sufficiently rapidly at spatial infinity. In order to use asymptotically PG slices, it is necessary to choose the conjugate to $M$ in terms of the PG time $T_{\rm PG}$. That is \be \tilde{P}_M = -e^\sigma T'_{\rm PG}= \frac{y\Lambda}{F} - \frac{\sqrt{1-F}}{FR'}. \label{eq:PM PG} \ee Since the extra term on the right is just a function of $R$ and $M$, this corresponds to a straightforward canonical transformation $(M,P_M,S,P_S)\to (M,\tilde{P}_M,S,\tilde{P}_S)$. For simplicity, we henceforth stick to the Schwarzschild expressions. From the expression (\ref{MS}) for the Misner-Sharp mass, we obtain \begin{align} P_M{\dot M} =&\frac{(n-2)\ma A_{n-2}}{2\kappa_n^2}\frac{y\Lambda}{F}R^{n-3}\biggl[{\dot F}-(n-3)(1-F)\frac{\dot R}{R}\biggl],\label{P_MdotMgr} \end{align} which shows $P_M{\dot M} =P_\Lambda{\dot \Lambda}+(\cdots){\dot R}+\delta(\cdots)+(\cdots)'$. The Misner-Sharp mass $M$ is expressed in terms of $\{\Lambda,P_\Lambda; R,P_R\}$ as \begin{align} M=&\frac{\kappa_n^2P_\Lambda^2}{2(n-2)\ma A_{n-2}R^{n-3}}+\frac{(n-2)\ma A_{n-2}}{2\kappa_n^2}R^{n-3}\biggl(1-\frac{{R'}^2}{\Lambda^2}\biggl).\label{MLiou} \end{align} From this expression, we can show that $M'$ is a linear combination of the constraints: \begin{align} M'=\Lambda^{-1}(yH_r-R'H),\label{M'2} \end{align} where we used Eq.~(\ref{PLambdaGR2}) to replace $P_\Lambda$ by $y$. This implies that in the vacuum theory $M$ is a constant on the constraint surface, as expected. Also using the expression (\ref{MLiou}), we can show that two sets of variables $\{\Lambda, P_\Lambda; R,P_R\}$ and $\{M, P_M; S,P_S\}$ satisfy the following Liouville form: \begin{align} P_\Lambda\delta \Lambda+P_R\delta R=P_M\delta M+P_S\delta S+\delta \eta+\zeta',\label{Liouville} \end{align} where \begin{align} \eta:=&\Lambda P_\Lambda+\frac{(n-2)\ma A_{n-2}}{2\kappa_n^2}R^{n-3}R' \ln\biggl|\frac{R'+y\Lambda}{R'-y\Lambda}\biggl|,\label{deta-gr}\\ \zeta:=&-\frac{(n-2)\ma A_{n-2}}{2\kappa_n^2}R^{n-3}\ln\biggl|\frac{R'+y\Lambda}{R'-y\Lambda}\biggl|\delta R. \label{zeta'-gr} \end{align} Under the boundary condition (\ref{bc1})--(\ref{bc6}), the total derivative term $\zeta$ converges to zero at spacelike infinity. Hence, the transformation from a set $\{\Lambda, P_\Lambda; R,P_R\}$ to $\{M,P_M;S,P_S\}$ is indeed a canonical transformation, namely \begin{align} &\int_{-\infty}^{\infty} dx(P_\Lambda\delta \Lambda+P_R\delta R)-\int_{-\infty}^{\infty} dx(P_M\delta M+P_S\delta S)=\delta \omega[\Lambda, P_\Lambda,; R,P_R], \label{canonical}\\ &\omega[\Lambda, P_\Lambda,; R,P_R]:=\int_{-\infty}^{\infty}dx\eta[\Lambda, P_\Lambda,; R,P_R] \end{align} is satisfied. It is shown that the integrands in the above equation, namely $P_\Lambda\delta \Lambda+P_R\delta R$, $P_M\delta M+P_S\delta S$, and $\eta$ converge to zero faster than $O(x^{-1})$ at spacelike infinity under the boundary condition we adopt, and hence the above expression is well-defined. (See Appendix~\ref{app:boundary} for the proof.) We now derive the Hamiltonian constraint and the diffeomorphism (momentum) constraint in terms of the variables $\{M,P_M;S,P_S\}$. A straightforward calculation using the above equations verifies the following relation; \begin{align} {\cal L}-P_M{\dot M}-\frac{N\Lambda}{R'}M'=&-\frac{{\dot R}}{R'}P_MM'+\mbox{(t.d.)} \\ =&\frac{y\Lambda}{F}\biggl(N_r+N\frac{y}{R'}\biggl)M'+\mbox{(t.d.)}, \label{eq:GRfinal} \end{align} where $\mbox{(t.d.)}$ is a total derivative term, we obtain the Hamiltonian density ${\cal H}_{\rm G}$ in the equivalent two-dimensional theory as \begin{align} {\cal H}_{\rm G}:=& P_M\dot{M} + P_S\dot{S}-{\cal L} \nonumber \\ =& N^M M'+N^S P_S, \label{eq:geometrodynamic hamiltonian} \end{align} where we have used Eq.~(\ref{eq:GRfinal}) and defined new Lagrange multipliers $N^M$ and $N^S$ as \begin{align} N^M:=&- \frac{\Lambda}{R'}\biggl(N+\frac{y{\dot R}}{F}\biggl) \nonumber\\ =&- \frac{\Lambda}{R'}\biggl(N+\frac{y}{F}\left(Ny+N_r R'\right)\biggl) \nonumber\\ =&-N\left(\frac{\Lambda}{R'}-\frac{P_M y}{R'}\right) +N_r P_M, \label{NM} \\ N^S:=&\dot{S}= \left(Ny+N_r R'\right).\label{NS} \end{align} Collecting terms in $N$ and $N_r$, we can express the total Hamiltonian as \begin{equation} {\cal H}_{\rm G}= N\left[\left(\frac{P_M y}{R'} -\frac{\Lambda}{R'}\right)M'+ y P_s\right] + N_r (P_M M' +P_S S'). \label{eq: HG} \end{equation} The coefficients of $N$ and $N_r$ are the super-Hamiltonian $H$ and the super-momentum $H_r$, respectively. These can in principle be expressed in terms of Kucha\v{r}' variables by doing the inverse canonical transformation. Note that one can replace $H$ by the linear combination of constraints \begin{equation} {\cal G} := H - \frac{y}{R'} H_r = -\frac{\Lambda}{R'} M' \end{equation} in agreement with (\ref{M'2}). The Lagrangian density for the canonical coordinates $(M,S,N^M,N^S)$ is now written as \begin{align} {\cal L}= P_M\dot{M} + P_S\dot{S}- N^M M'-N^S P_S, \label{eq:geometrodynamic L} \end{align} which corresponds to Eq.~(122) of \cite{kuchar94}. The constraints, $M'=0$ and $P_S=0$, are obtained by varying the Lagrange multipliers $N^M$ and $N^S$, respectively. On the constraint surface $M=m(t)$ and $P_S=0$ hold. The reduced phase space is therefore two-dimensional consisting of $p_m:=\int^\infty_{-\infty} dx P_M(x,t)$ and $m$. With suitable boundary conditions~\cite{KMT12}, one can repeat the analysis of~\cite{kuchar94} for spacelike slicings that intersect both left and right branches of the outer horizons of eternal black holes to obtain the reduced action: \begin{align} I_{(2)} = \int dt \biggl[p_m\dot{m} - ({N}_+-{N}_-)m\biggl], \label{eq:reduced action} \end{align} where $N_{\pm}:=\mp \lim_{x\to \pm\infty}N^M$. The reduced equations of motion in vacuum then imply that $m=m_0=constant$, and $\dot{p}_m = -(N_+-N_-)$. Lastly, let us derive the boundary term in Eq.~(\ref{boundaryGR}) with the new canonical variables. Starting from \begin{align} I_{\rm M(GR)}=\int dt\int dx(P_M\dot{M} + P_S\dot{S}- N^M M'-N^S P_S), \end{align} we obtain \begin{align} \delta I_{\rm M(GR)}=&\int dt\int dx\biggl(\delta P_M\dot{M}+\partial_t(P_M\delta M)-{\dot P}_M\delta {M}+ \delta P_S\dot{S} + \partial_t(P_S\delta S)-{\dot P}_S\delta S \nonumber \\ &- \delta N^M M'- (N^M \delta M)'+{N^M}' \delta M-\delta N^S P_S-N^S \delta P_S\biggl) \nonumber \\ =&\int dt\int dx \biggl(\mbox{dynamical terms}\biggl)+\int dx \biggl[P_M\delta M+P_S\delta S\biggl]_{t=t_1}^{t=t_2}-\int dt\biggl[N^M \delta M\biggl]_{x=-\infty}^{x=+\infty}. \end{align} Under the boundary condition (\ref{bc1})--(\ref{bc6}), we obtain $\delta M \simeq \delta M^\infty(t)$ and $N^M\simeq -N_\infty(t)$ at spacelike infinity and hence we obtain the same result (\ref{boundaryGR}) by setting $\delta M=0$ and $\delta S=0$ at $t=t_1,t_2$. One important advantage of the new set of canonical variables is to greatly simplify the calculations. We will take advantage of this simplification in the next section. \section{Canonical formalism in Lovelock gravity} \label{Can form Lovelock} In this section, we show that all the results in the previous section can be generalized to full Lovelock gravity. In particular the transformation from the ADM variables $\{\Lambda, P_\Lambda; R,P_R\}$ to $\{M, P_M; S,P_S\}$ is a well-defined canonical transformation using definitions of $P_M$, $S$, and $P_S$ that are the same as those in general relativity, Eqs.~(\ref{eq:R})--(\ref{eq:Pr1}), and $M$ defined by Eq.~(\ref{qlm-L}). \subsection{ADM variables} First we derive the ADM conjugate momenta $P_\Lambda$ and $P_R$. The Lagrangian density from the action (\ref{action-3}) is \begin{align} {\cal L}=& \frac{(n-2){\cal A}_{n-2}}{2\kappa_n^2}\sum^{[n/2]}_{p=0}{\tilde \alpha}_{(p)} \Biggl[2pR^{n-2p-1}y(N_r\Lambda)' -\frac{2pN}{n-2p}\biggl((R^{n-2p})'\Lambda^{-1}\biggl)' \nonumber \\ & + (n-2p-1)\biggl\{\left(1-F\right)^{p} +2pF\biggl\}N\Lambda R^{n-2-2p}-2pR^{n-2p-1}y {\dot \Lambda} \nonumber \\ &+ pR^{n-2p-1}\biggl\{1-(1-F)^{p-1}\biggl\}\biggl\{(\Lambda N_r y+\Lambda^{-1}NR')\frac{F'}{F}-\Lambda y\frac{\dot F}{F} \biggl\}\biggl].\label{Lag-0} \end{align} Using the binomial expansion and integration by parts many times, we can rewrite the above Lagrangian density into the following form up to the total derivative. The derivation is presented in Appendix~\ref{appendix2}. \begin{align} {\cal L}=& \frac{(n-2){\cal A}_{n-2}}{2\kappa_n^2}\sum^{[n/2]}_{p=0}{\tilde \alpha}_{(p)} \Biggl[2pR^{n-2p-1}y(N_r\Lambda)'-\frac{2pN}{n-2p}\biggl((R^{n-2p})'\Lambda^{-1}\biggl)' \nonumber \\ & + (n-2p-1)\biggl\{\left(1-F\right)^{p} +2pF\biggl\}N\Lambda R^{n-2-2p}-2pR^{n-2p-1}y {\dot \Lambda}\biggl] \nonumber \\ &-\frac{(n-2){\cal A}_{n-2}}{2\kappa_n^2}\sum^{[n/2]}_{p=2}{\tilde \alpha}_{(p)}\biggl[\sum_{w=0}^{p-2}\frac{p!(-1)^{p-1-w}}{w!(p-1-w)!}\biggl\{ \Lambda N_r yR^{n-2p-1}F^{p-2-w}(\Lambda^{-2}{R'}^2)' \nonumber \\ &+\sum_{j=0}^{p-2-w}\frac{2(p-2-w)!(-1)^{p-2-w-j}}{j!(p-2-w-j)!}\frac{(N_r R^{n-2p-1}\Lambda^{1-2j}{R'}^{2j})'y^{2(p-w-j)-1}}{2(p-w-j)-1} \nonumber \\ &-\frac{(\Lambda^{-1}NR'R^{n-2p-1})'F^{p-1-w}}{p-1-w}\biggl\} +\sum_{w=1}^{p-1}\frac{2p!(-1)^w}{w!(p-1-w)!}R^{n-2p-1} F^{w-1} y\Lambda^{-2}{\dot \Lambda}{R'}^2\nonumber \\ &-\sum_{w=1}^{p-1}\frac{2p!}{w(p-1-w)!}\sum_{j=0}^{w-1}\frac{(-1)^{2w-1-j}}{j!(w-1-j)!} \biggl\{\frac{\partial_t(R^{n-2p-1}\Lambda^{1-2j}){R'}^{2j}y^{2(w-j)+1} }{2(w-j)+1}\nonumber \\ &-j\sum_{q=0}^{2(w-j)+1}\frac{(2w-2j)!(-1)^q}{q!(2w-2j+1-q)!}\frac{{\dot R}^{w-j+1}(R^{n-2p-1}\Lambda^{1-2j}N^{-2(w-j)-1}{R'}^{2j-1+q}N_r^q)'}{2(w-j+1)} \nonumber \\ &-\sum_{q=0}^{2(w-j)-1}\frac{(2w-2j-1)!(-1)^q}{q!(2w-2j-q)!}{\dot R}^{2w-2j-q}(R^{n-2p-1}\Lambda^{-1-2j}N^{-2(w-j)+1}{R'}^{2j+1+q}N_r^q)'\biggl\}\biggl] \label{Lag-1}. \end{align} From this Lagrangian density, we obtain \begin{align} P_\Lambda=& -\frac{(n-2){\cal A}_{n-2}}{\kappa_n^2}\biggl[\sum^{[n/2]}_{p=0}{\tilde \alpha}_{(p)}pR^{n-2p-1}y +\sum^{[n/2]}_{p=2}{\tilde \alpha}_{(p)}R^{n-2p-1}y\frac{{R'}^2}{\Lambda^{2}} \nonumber \\ &\times \sum_{w=1}^{p-1}\frac{p!(-1)^w}{w!(p-1-w)!}\biggl\{F^{w-1}- \sum_{j=0}^{w-1}\frac{(-1)^{w-1-j}}{j!(w-1-j)!} \frac{(1-2j)y^{2(w-j)}}{2(w-j)+1}\frac{{R'}^{2j-2}}{\Lambda^{2j-2}}\biggl\} \biggl]\label{PLambda3} \end{align} and \begin{align} P_R=& \frac{(n-2){\cal A}_{n-2}}{\kappa_n^2}\sum^{[n/2]}_{p=0}{\tilde \alpha}_{(p)} pR^{n-2p-1} \Biggl[\frac{(N_r\Lambda)'-{\dot \Lambda}}{N}+(n-2p-1)\biggl\{\left(1-F\right)^{p-1} -2\biggl\}\frac{y\Lambda}{R}\biggl] \nonumber \\ &-\frac{(n-2){\cal A}_{n-2}}{2\kappa_n^2}\sum^{[n/2]}_{p=2}{\tilde \alpha}_{(p)} p\biggl[\sum_{w=0}^{p-2}\frac{(p-1)!(-1)^{p-1-w}}{w!(p-1-w)!}N^{-1}\biggl\{2(\Lambda^{-1}NR'R^{n-2p-1})'F^{p-2-w}y\nonumber \\ &+ \sum_{j=0}^{p-2-w}\frac{2(p-2-w)!(-1)^{p-2-w-j}}{j!(p-2-w-j)!}(N_r R^{n-2p-1}\Lambda^{1-2j}{R'}^{2j})'y^{2(p-w-j-1)} \nonumber \\ &+\Lambda N_r R^{n-2p-1}F^{p-3-w}(\Lambda^{-2}{R'}^2)'\biggl(F-2(p-2-w) y^2\biggl) \biggl\} \nonumber \\ &+\sum_{w=1}^{p-1}\frac{2(p-1)!(-1)^w}{w!(p-1-w)!}\biggl\{R^{n-2p-1} \Lambda^{-2}{\dot \Lambda}{R'}^2N^{-1}F^{w-2}\biggl(F-2(w-1) y^2\biggl)\nonumber \\ &- \sum_{j=0}^{w-1}\frac{(-1)^{w-1-j}}{j!(w-1-j)!}\biggl(\partial_t(R^{n-2p-1}\Lambda^{1-2j})\frac{{R'}^{2j}y^{2(w-j)}}{N} +\frac{n-2p-1}{2(w-j)+1}\frac{R^{n-2p-2}{R'}^{2j}y^{2(w-j)+1}}{\Lambda^{2j-1}} \nonumber \\ &-\sum_{q=0}^{2(w-j)+1}\frac{2j(2w-2j)!(-1)^q}{q!(2w-2j+1-q)!}{\dot R}^{2(w-j)+1}(R^{n-2p-1}\Lambda^{1-2j}N^{-2(w-j)-1}{R'}^{2j-1+q}N_r^q)' \nonumber \\ &-\sum_{q=0}^{2(w-j)-1}\frac{(2w-2j-1)!(-1)^q}{q!(2w-2j-1-q)!}{\dot R}^{2w-2j-q-1}(R^{n-2p-1}\Lambda^{-1-2j}N^{-2(w-j)+1}{R'}^{2j+1+q}N_r^q)'\biggl)\biggl\}\biggl]. \end{align} In general relativity, $P_\Lambda=P_\Lambda[{\dot \Lambda},y({\dot R})]$ and $P_R=P_R[{\dot \Lambda},y({\dot R})]$ can be algebraically solved to give a unique set of ${\dot \Lambda}={\dot \Lambda}[P_\Lambda,P_R]$ and ${\dot R}={\dot R}[P_\Lambda,P_R]$. In higher-order Lovelock gravity, by contrast, it is not possible to obtain a unique expression in general because of the fact that $P_\Lambda=P_\Lambda[{\dot \Lambda},y({\dot R})]$ and $P_R=P_R[{\dot \Lambda},y({\dot R})]$ are higher-order polynomials of $y$. As a result, it is difficult to obtain the explicit forms of the super-momentum $H_r$ and the super-Hamiltonian $H$, such that \begin{align} {\cal L}= {\dot \Lambda}P_{\Lambda}+{\dot R}P_{R}-N H-N_r H_{r} \end{align} in terms of the ADM variables. However, it is not necessary to do so at this stage. Things are greatly simplified by using the generalized Misner-Sharp mass as a new canonical variable. As we will show, the super-momentum and the super-Hamiltonian with the new set of canonical coordinates are the same as those in general relativity and then the boundary terms at spatial infinity can be easily derived. \subsection{Generalized Misner-Sharp mass as canonical variable} We introduce a new set of canonical variables $\{M,P_M;S, P_S\}$ defined in the same way as in general relativity, namely by Eq.~(\ref{qlm-L}), and Eqs.~(\ref{eq:R})--(\ref{eq:Pr1}). Then, we prove that $\{\Lambda,P_\Lambda;R, P_R\}$ and $\{M,P_M;S, P_S\}$ again satisfy the Liouville form (\ref{Liouville}) with the following total variation and the total derivative terms. The derivation is presented in Appendix~\ref{app:Liouville}. \begin{align} \eta:=&\frac{(n-2){\cal A}_{n-2}}{2\kappa_n^2}\biggl[\sum_{p=1}^{[n/2]}{\tilde \alpha}_{(p)}pR^{n-1-2p}\biggl(2y\Lambda-R' \ln\biggl|\frac{R'+y\Lambda}{R'-y\Lambda}\biggl|\biggl) \nonumber \\ &-\sum_{p=2}^{[n/2]}{\tilde \alpha}_{(p)}R^{n-1-2p} \sum_{w=1}^{p-1}\frac{p!}{w(p-1-w)!}\sum_{j=0}^{w-1}\frac{2(-1)^{2w-1-j}{R'}^{2j}y^{2(w-j)+1}}{j!(w-1-j)![2(w-j)+1]\Lambda^{2j-1}}\biggl], \label{deta} \\ \zeta:=&\frac{(n-2){\cal A}_{n-2}}{2\kappa_n^2}\biggl[\sum_{p=1}^{[n/2]}{\tilde \alpha}_{(p)}pR^{n-1-2p}\ln\biggl|\frac{R'+y\Lambda}{R'-y\Lambda}\biggl| +\sum_{p=2}^{[n/2]}{\tilde \alpha}_{(p)}R^{n-1-2p}\nonumber \\ &\times \sum_{w=1}^{p-1}\frac{2p!(-1)^wyR'}{w!(p-1-w)!\Lambda}\biggl\{F^{w-1}+\sum_{j=0}^{w-1}\frac{2j(w-1)!(-1)^{w-1-j}{R'}^{2j-2}y^{2(w-j)}}{j!(w-1-j)![2(w-j)+1]\Lambda^{2j-2}}\biggl\}\biggl]\delta R. \label{zeta'} \end{align} In Appendix~\ref{appendix3}, it is proven that Eq.~(\ref{eq:GRfinal}) still holds in full Lovelock gravity. This immediately implies that the Hamiltonian density in the equivalent two-dimensional theory takes the same form as that in general relativity (\ref{eq:geometrodynamic hamiltonian}), where the definitions of the new Lagrange multipliers $N^M$ and $N^S$ are the same as those in general relativity (\ref{NM}) and (\ref{NS}). Finally, the Lagrangian density for the canonical coordinates $\{M,P_M;S,P_S\}$ can be again written as \begin{align} {\cal L}= P_M\dot{M} + P_S\dot{S}- N^M M'-N^S P_S \label{eq:geometrodynamic L2} \end{align} and the super-Hamiltonian and super-momentum constraints are again as in (\ref{eq: HG}). In comparison to the rather complicated starting point in Eq.~(\ref{L_p}), this equivalent Lagrangian density is extremely simple and the physical meaning of the canonical variables are very clear. Remarkably, the coupling constants $\alpha_{(p)}$ do not appear explicitly in (\ref{eq:geometrodynamic L2}). They are in fact hidden in the definition of the mass function. This makes it possible to treat any class of Lovelock gravity in exactly the same way. \subsection{Fall-off rate at infinity and boundary terms} In order to prove that the transformation from $\{\Lambda, P_\Lambda,; R,P_R\}$ to $\{M,P_M;S,P_S\}$ is canonical and well-defined, we have to discuss the asymptotic behaviour of the variables. We adopt the same boundary conditions (\ref{bc1})--(\ref{bc4}) as in general relativity. With these conditions, one can verify that the generalized Misner-Sharp mass (\ref{qlm-L}) behaves near spacelike infinity as \begin{align} M \simeq \frac{(n-2)A_{n-2}{\tilde \alpha}_{(1)}\Lambda_1(t)}{\kappa_n^2}. \end{align} This is the same as in general relativity and hence we set $\Lambda_1$ as in Eq.~(\ref{bc7}) (where ${\tilde \alpha}_{(1)}=1$) in order that $M\simeq M^\infty(t)$ at infinity. It can then be shown that the leading terms of $P_\Lambda$, $P_R$, $\zeta$, $\eta$, $P_S$, $P_M$, $N^M$, and $N^S$ are the same as those in the general relativistic case under the boundary condition (\ref{bc1})--(\ref{bc4}). As a consequence, the proof carries over from general relativity and all the terms in the Liouville form (\ref{Liouville}) are well behaved at spacelike infinity. This is sufficient to prove the transformation from $\{\Lambda, P_\Lambda,; R,P_R\}$ to $\{M,P_M;S,P_S\}$ is indeed a well-defined canonical transformation. and that the Hamiltonian $\int^\infty_{-\infty}dx(N^M M'+N^S P_S)$ is also finite. We now have the following two-dimensional action with a new set of canonical variables; \begin{align} I_{\rm M(L)}=\int dt\int dx(P_M\dot{M} + P_S\dot{S}- N^M M'-N^S P_S), \end{align} with the same asymptotic behavior as in general relativity. The boundary term for the above action that makes the variational principle well defined is then also the same as in general relativity: \begin{align} \delta I_{\rm M(L)}=\int dt\int dx \biggl(\mbox{dynamical terms}\biggl)+\int dt\biggl[N_\infty(t)\delta M^\infty(t)\biggl]_{x=-\infty}^{x=+\infty}. \end{align} Given the above, we can now write down the super-Hamiltonian and super-momentum constraints for full Lovelock gravity. In terms of the geometrodynamical variables they are the same expressions as in general relativity: \bea H &=& \left(\frac{P_M y}{R'} -\frac{\Lambda}{R'}\right)M' + y P_s\, , \label{eq:super H} \\ H_r &=& P_M M' +P_S S'. \label{eq:super Hr} \eea The expressions in terms of ADM variables are considerably more complicated and can in principle be obtained once again by substution from Eqs.~(\ref{eq:R})--(\ref{eq:Pr1}), with $M$ defined by Eq.~(\ref{qlm-L}). \section{Adding matter fields} \label{matter} In this section, we introduce matter fields in the argument with the ADM variables discussed in the previous sections. Here we write super-momentum and super-Hamiltonian for gravity as $H_r^{\rm (G)}$ and $H^{\rm (G)}$ in order to distinguish from the total super-momentum and super-Hamiltonian including matter contributions. The following argument is valid in full Lovelock gravity. It can be shown from Eqs.~(\ref{defy}), (\ref{eq:PM}), (\ref{eq:R}), (\ref{eq:Pr1}) and (\ref{eq:GRfinal}) that the gravitational Hamiltonian $H_{\rm G}$ is given by: \be \label{HGADM} H_{\rm G} = \int dx (NH^{\rm (G)} + N_rH_r^{\rm (G)}), \ee where \begin{align} \label{momconst1} H_r^{\rm (G)} =& P_S S^\prime + P_M M^\prime = P_R R^\prime - P_\Lambda^\prime \Lambda,\\ \label{Hconst1} H^{\rm (G)} =& -\frac{\Lambda}{R^\prime}M^\prime + \frac{y}{R^\prime}H_r^{(G)}. \end{align} We have used (\ref{M'2}) to derive (\ref{Hconst1}). Since Eq.~(\ref{PLambda3}) shows that $y$ is not a function of $N$ or $N_r$, we can see that the Hamiltonian density is the sum of Lagrange multiplier times constraints. \subsection{Massless scalar field} \label{Scalar field} First we consider a massless scalar field $\psi$ as a matter field, of which action is $I_{\rm matter}=I_\psi$ in the action (\ref{action}): \begin{align} I_\psi = -\frac{1}{2}\int d^{n}x \sqrt{-g} (\nabla \psi)^2. \end{align} The equivalent two-dimensional action in the symmetric spacetime under consideration is given by \begin{align} \label{scalaraction2} I_\psi &= -\frac{{\cal A}_{n-2}}{2}\int d^2{\bar y} \sqrt{-\g} R^{n-2} (D \psi)^2 \\ \nonumber & = -\frac{{\cal A}_{n-2}}{2}\int dxdt \frac{\Lambda R^{n-2} }{N} \left( -\psid^2 + 2N_r\psip \psid + (N^2\Lambda^{-2}-N_r^2) \psip^2 \right). \end{align} This gives the momentum conjugate $\Pis$ to $\psi$ as \be \label{Pip} \Pis = \frac{{\cal A}_{n-2}\Lambda R^{n-2}}{N} \left( \psid - N_r \psip \right), \ee with which we can write the matter action as \begin{align} \label{scalaraction3} I_\psi=& \int dxdt \psid \Pis- \int dxdt N \left[ \frac{1}{2\Lambda} \left( \frac{\Pis^2}{{\cal A}_{n-2} R^{n-2}} +{\cal A}_{n-2}R^{n-2} \psip^2 \right) + \Pis \psip\frac{N_r}{N} \right]. \end{align} Equation (\ref{scalaraction2}) does not contain any derivatives of the metric or $R$, which means that adding the scalar action to the gravitational action (\ref{eq:geometrodynamic hamiltonian}) does not change $P_\Lambda$ or $P_R$. This allows us to write the total Hamiltonian as the sum of the gravitational and matter parts. Using Eqs.~(\ref{HGADM}), (\ref{momconst1}), (\ref{Hconst1}), and (\ref{scalaraction3}), we obtain the total Hamiltonian $H_{\rm total} $ as \begin{align} \label{H8} H_{\rm total} & = \int dx N\Bigg[-\frac{\Lambda}{R^\prime}M^\prime + \frac{y}{R^\prime}H_r + \frac{N_r}{N} H_r + \frac{1}{2\Lambda} \left( \frac{P_\psi^2}{{\cal A}_{n-2} R^{n-2}} + {\cal A}_{n-2} R^{n-2} \psip^2 \right) +P_\psi \psip\frac{N_r}{N} \Bigg] \none & = \int dx \biggl[ N(H^{\rm (G)}+H^{\rm (M)}) + N_r(H_r^{\rm (G)}+H_r^{\rm (M)}) \biggl], \end{align} where $y$ is a function of the phase space variables, $\Lambda$, $P_\Lambda$ and $R$ via Eq.~(\ref{PLambda3}). The super-Hamiltonian $H^{\rm (M)}$ and super-momentum $H_r^{\rm (M)}$ for $\psi$ are given by \begin{align} \label{HconstM} H^{\rm (M)} =& \frac{1}{2\Lambda} \left( \frac{P_\psi^2}{{\cal A}_{n-2} R^{n-2}} +{\cal A}_{n-2}R^{n-2} \psip^2 \right),\\ \label{momconstM} H_r^{\rm (M)} =& P_\psi \psip. \end{align} The Poisson bracket of Hamiltonian constraint, $H = H^{\rm (G)} + H^{\rm (M)}$ with the total momentum constraint, $H_r = H_r^{\rm (G)} + H_r^{\rm (M)}$ is given by \be \label{firstclasstotal} \{H, H_r \} = \{H^{\rm (G)}, H_r^{\rm (G)} \} + \{H^{\rm (G)}, H_r^{\rm (M)} \} + \{H^{\rm (M)}, H_r^{\rm (G)} \} + \{H^{\rm (M)}, H_r^{\rm (M)} \}. \ee Because our theory is diffeomorphism invariant, $\{H, H_r \}$ must be weekly equal to zero. Because there are two first class constraints, there are two gauge choices to pick. We choose our first gauge as \be \label{gauge1} \chi := R - x \approx 0. \ee This forces the spatial coordinate to be the areal radius which means that $R$ is no longer a phase space variable, it is now a coordinate. In order to insist that $\chi$ is satisfied at every time slice, we must insist that $\dot{\chi}=\{\chi,H\} \approx 0$, which shows $N_r/N +y/R^\prime \approx 0$. We use this relation to write one Lagrange multiplier in terms of the other. This leaves us with one Lagrange multiplier which reflects the fact that there is only one gauge fix left to choose. We can now plug the gauge choice (\ref{gauge1}) and its consistency condition into the Hamiltonian as long as we use Dirac brackets to evaluate the equations of motion in the end. Note that the remaining phase space variables, $\Lambda$, $P_\Lambda$, $\psi$ and $P_\psi$, all commute with $\chi$ and so the Poisson bracket is the same as the Dirac bracket. Plugging $\chi = \dot{\chi} = 0$ into Eq.~(\ref{H8}) gives \be \label{H8b} H_{\rm total} = \int dR N \Biggl[-\Lambda M^\prime + \frac{1}{2\Lambda} \left( \frac{\Pis^2}{{\cal A}_{n-2}R^{n-2}} + {\cal A}_{n-2}R^{n-2} \psip^2 \right) - y \Pis \psip \Biggr]. \ee In the last term we replaced $N_r/N$ by $-y$ as required. Since the mass equation (\ref{qlm-L}) is written as \begin{align} \label{Massfunc} M=\frac{(n-2){\cal A}_{n-2}}{2\kappa_n^2}\sum_{p=0}^{[n/2]} \tilde{\alpha}_{(p)}R^{n-1-2p}\left(1-\Lambda^{-2} + y^2\right)^p, \end{align} we can write $y$($=N_r/N$) in terms of the mass function. For this reason we leave the factor of $N_r/N$ in the Hamiltonian with the understanding that it is the solution to Eq.~(\ref{Massfunc}). For our second gauge choice we choose \be \label{gauge2} \xi := \Lambda - 1\approx 0. \ee By the same reasoning used for the first gauge choice we can set $\xi$ strongly to zero (namely since $\Lambda$ commutes with $\psi$ and $P_\psi$) which gives the Hamiltonian \be \label{H8d} H_{\rm total} = \int dR N \Biggl[ -M^\prime + \frac{1}{2} \left( \frac{\Pis^2}{{\cal A}_{n-2} R^{n-2}} + {\cal A}_{n-2} R^{n-2} \psip^2 \right) + \Pis \psip \frac{N_r}{N} \Biggr] \ee and the mass function \begin{align} \label{Nsig} M= \frac{(n-2){\cal A}_{n-2}}{2\kappa_n^2}\sum_{p=0}^{[n/2]} \tilde{\alpha}_{(p)}R^{n-1-2p}\left(\frac{N_r}{N}\right)^{2p}. \end{align} To see the significance of this gauge choice, notice from Eq.~(\ref{qlm-L}) that $g^{11} \to 1-2\kappa_n^2M/[(n-2){\ma A_{(n-2)}}\tilde{\alpha}_{(1)}R^{n-3}]$ in the general relativistic case when we strongly set $\xi$ and $\chi$ to zero. This gives the metric in the non-static version of Painlev\'{e}-Gullstrand coordinates: \be ds_{(2)}^2= -N^2 \left( 1-\frac{2\kappa_n^2M}{(n-2){\ma A_{(n-2)}}\tilde{\alpha}_{(1)}R^{n-3}} \right) dt^2 + 2 N \sqrt{\frac{2\kappa_n^2M}{(n-2){\ma A_{(n-2)}}\tilde{\alpha}_{(1)}R^{n-3}}} dtdR + dR^2. \ee To ensure that the second gauge condition is conserved in time we must insist that $d(\Lambda - 1)/dt = \{\Lambda - 1, H\} = 0 \rightarrow \delta H/\delta P_\Lambda = 0$. Although we have chosen to write $N_r/N$ in terms of the mass function, it can also be written in terms of $P_\Lambda$. All of the $P_\Lambda$ dependence in the Hamiltonian is in the terms of $N_r/N$. Therefore we can write \begin{align} \frac{\delta H_{\rm total}}{\delta P_\Lambda} & = \frac{\delta}{\delta P_\Lambda} \int dR \left( N^\prime M + N P_\psi \psip \frac{N_r}{N} \right) \nonumber \\ & = N^\prime \frac{\partial M}{\partial (N_r/N)} \frac{\partial (N_r/N)}{\partial P_\Lambda} + N P_\psi \psip\frac{\partial (N_r/N)}{\partial P_\Lambda}, \label{consistency conditon 2} \end{align} from which the consistency condition is given as \be \label{sigma1} N^\prime \frac{\partial M}{\partial (N_r/N)} + N P_\psi \psip = 0, \ee where it is understood that we use (\ref{Nsig}) to find $\partial M/\partial (N_r/N)$ and write $N_r/N$ in terms of the mass function $M$. Notice that the actual relation between $N_r/N$ and $P_\Lambda$ is not needed. Using Hamilton's equations and Eq.~(\ref{H8d}), we find \begin{align} \label{psidot} \dot{\psi} =& N \left( \frac{\Pis}{{\cal A}_{n-2} R^{n-2}} + \psip\frac{N_r}{N} \right),\\ \label{Pisdot} \dot{P}_\psi =& \left[ N \left( {\cal A}_{n-2} R^{n-2} \psip + \Pis \frac{N_r}{N} \right) \right]^\prime. \end{align} These equations, along with the consistency conditions (\ref{Nsig}) and (\ref{sigma1}) and the Hamiltonian constraint \be \label{Cconstraint} -M^\prime + \frac{1}{2} \left( \frac{\Pis^2}{{\cal A}_{n-2}R^{n-2}} + {\cal A}_{n-2} R^{n-2} \psip^2 \right) + \Pis \psip \frac{N_r}{N}= 0, \ee determine the evolution of a collapsing scalar field. \subsection{Charged scalar field} In this subsection, we consider a U(1) gauge field $A_\mu$ coupled to a charged complex massless scalar field $\psi = (\psi_1 + i\psi_2)/\sqrt{2}$, where $\psi_1$ and $\psi_2$ are real functions. We write the action for this matter as $I_{\rm matter}=I_{\rm EM}$: \be \label{EMaction1} I_{\rm EM}= \int d^nx \sqrt{-g}\left[-\left( \partial^\mu + ieA^\mu \right) \psi^* \left( \partial_\mu - ieA_\mu \right) \psi- \frac{1}{4} F^{\mu \nu} F_{\mu \nu} \right], \ee where $e$ is the charge and the Faraday tensor $F_{\mu \nu}$ is defined in terms of the gauge field as $F_{\mu \nu}=\partial_\mu A_\nu-\partial_\nu A_\mu$. Under the symmetry assumption in the present paper both for gravity and matter, the equivalent two-dimensional action is given by \be \label{EMaction2} I_{\rm EM}={\cal A}_{n-2}\int dtdx \sqrt{-g_{(2)}} R^{n-2}\left[-\left( \partial^B + ieA^B \right) \psi^* \left( \partial_B - ieA_B \right) \psi -\frac{1}{4} F^{AB} F_{AB}\right]. \ee Adopting the ADM coordinates, we obtain \begin{align} \label{EMaction3} I_{\rm EM} =&\frac{{\cal A}_{n-2}}{2} \int dtdx \frac{R^{n-2}\Lambda}{N}\Bigg[(\dot{\psi}_1^2 + \dot{\psi}_2^2) - 2N_r(\dot{\psi_1}\psi_1^\prime + \dot{\psi_2}\psi_2^\prime) +(N_r^2-N^2\Lambda^{-2})(\psi_1^{\prime 2} + \psi_2^{\prime 2}) \nonumber \\ & -2e \biggl\{ (A_0-N_rA_1)(\dot{\psi}_2 \psi_1 - \dot{\psi}_1 \psi_2) - \biggl(N_r(A_0-N_rA_1) + N^2\Lambda^{-2}A_1\biggl)(\psi_2^\prime \psi_1 - \psi_1^\prime \psi_2) \biggl\}\nonumber \\ & +e^2 \biggl((A_0-N_rA_1)^2-N^2\Lambda^{-2}A_1^2\biggl)(\psi_1^2 + \psi_2^2) + \Lambda^{-2} (\dot{A_1} - A_0^\prime)^2 \Bigg], \end{align} where $A_\mu dx^\mu=A_0(t,x)dt+A_1(t,x)dx$. From the above action we find the conjugate momenta: \begin{align} \label{Ppsi1} P_{\psi 1} =& \frac{{\cal A}_{n-2}R^{n-2}\Lambda}{N}\left[ \dot{\psi}_1 - N_r \psi_1^\prime +e(A_0 - N_r A_1) \psi_2 \right],\\ \label{Ppsi2} P_{\psi 2} =&\frac{{\cal A}_{n-2}R^{n-2}\Lambda}{N}\left[ \dot{\psi}_2 - N_r \psi_2^\prime -e(A_0 - N_r A_1) \psi_1 \right],\\ \label{PA0} P_{A0} =& 0,\\ \label{PA1} P_{A1} =& \frac{{\cal A}_{n-2}R^{n-2}(\dot{A_1} - A_0^\prime)}{N\Lambda}, \end{align} which give the Hamiltonian for the present matter field: \begin{align} \label{HEM1} H_{\rm EM}=&\int dx \Bigg[ \frac{N}{2{\cal A}_{n-2}\Lambda R^{n-2}}(P_{\psi 1}^2 + P_{\psi 2}^2) + e(A_0-N_rA_1)(P_{\psi 2} \psi_1 - P_{\psi 1} \psi_2) \nonumber \\ & + N_r(P_{\psi 1} \psi_1^\prime + P_{\psi 2} \psi_2^\prime) + \frac{NR^{n-2}}{2{\cal A}_{n-2}\Lambda}\biggl\{ (eA_1\psi_1-\psi_2^\prime)^2+(eA_1\psi_2+\psi_1^\prime)^2 \biggl\} \nonumber \\ & + \frac{N\Lambda}{2{\cal A}_{n-2}R^{n-2}}P_{A1}^2 + P_{A1}A_0^\prime \Bigg]. \end{align} Since the action (\ref{EMaction3}) contains no derivatives of the metric or $R$, the addition of $I_{\rm EM}$ to the gravitational action does not alter the Hamiltonian analysis and allows us to write the total Hamiltonian as \be \label{HEMtotal1} H_{\rm total} = \int dx \biggl[N(H^{\rm (G)} + H^{\rm (EM)}) + N_r(H_r^{\rm (G)} + H_r^{\rm (EM)}) + A_0H_{A0}^{\rm (EM)} \biggl], \ee where $H^{\rm (G)}$, $H_r^{\rm (G)}$ and $H_{A0}^{\rm (EM)}$ are given by Eqs.~(\ref{Hconst1}) and (\ref{momconst1}) and $H^{\rm (EM)}$ and $H_r^{\rm (EM)}$ are given by \begin{align} H^{\rm (EM)}=& \frac{P_{\psi 1}^2 + P_{\psi 2}^2}{2{\cal A}_{n-2}\Lambda R^{n-2}} +\frac{{\cal A}_{n-2}R^{n-2}}{2\Lambda}\biggl[ (eA_1\psi_1-\psi_2^\prime)^2+(eA_1\psi_2+\psi_1^\prime)^2 \biggl]+ \frac{\Lambda P_{A1}^2}{2 {\cal A}_{n-2}R^{n-2}},\label{HEM}\\ \label{HrEM} H_r^{\rm (EM)}=& -eA_1(P_{\psi 2} \psi_1 - P_{\psi 1} \psi_2) + (P_{\psi 1} \psi_1^\prime + P_{\psi 2} \psi_2^\prime),\\ \label{HA0EM} H_{A0}^{\rm (EM)}=&e(P_{\psi 2} \psi_1 - P_{\psi 1} \psi_2)-P_{A1}^\prime, \end{align} where we used integration by parts and asymptotic condition $P_ {A1}A_0\to 0$ at infinity to derive Eq.~(\ref{HA0EM}). The consistency condition on the constraint (\ref{PA0}) is $\{P_{A0},H_{\rm total}\} = 0$, which gives $e(P_{\psi 2} \psi_1 - P_{\psi 1} \psi_2)-P_{A1}^\prime=0$. This condition is already added into the Hamiltonian with $A_0$ as its Lagrange multiplier. Since $P_{A0}$ is weekly equal to zero we can use the equation of motion for $P_{A0}$ to show that $H_{A0}^{\rm (EM)}$ is weekly equal to zero and is, therefore, a constraint in the same way as the constraints multiplying $N$ and $N_r$. This Hamiltonian is composed of three first class constraints which means that there are three gauge choices to make. Our first two gauge choices will be the same as in section~\ref{Scalar field}. Using similar reasoning we can write the Hamiltonian as \begin{align} H_{\rm total} =& \int dR \biggl[ N\biggl\{ -M^\prime +\frac{P_{\psi 1}^2 + P_{\psi 2}^2}{2{\cal A}_{n-2}R^{n-2}} +\frac{{\cal A}_{n-2}R^{n-2}}{2}\biggl( (eA_1\psi_1-\psi_2^\prime)^2 + (eA_1\psi_2+\psi_1^\prime)^2 \biggl) \nonumber \\ & +\frac{P_{A1}^2}{2{\cal A}_{n-2}R^{n-2}} + \frac{N_r}{N} \biggl( -eA_1(P_{\psi 2} \psi_1 - P_{\psi 1} \psi_2) + (P_{\psi 1} \psi_1^\prime + P_{\psi 2} \psi_2^\prime) \biggl) \biggl\} \nonumber \\ & + A_0\biggl(e(P_{\psi 2} \psi_1 - P_{\psi 1} \psi_2)-P_{A1}^\prime\biggl)\biggl]. \label{HEMtotal1b} \end{align} Just as in section \ref{Scalar field} the consistency condition on the first gauge fix requires us to write $N_r/N$ as a function of $M$ using Eq.~(\ref{Nsig}). The consistency condition on the second gauge choice, analogous to Eq.~(\ref{sigma1}), is given by \be \label{EM consistency condition 2} N^\prime \frac{\partial M}{\partial (N_r/N)} + N \biggl( -eA_1(P_{\psi 2} \psi_1 - P_{\psi 1} \psi_2) + (P_{\psi 1} \psi_1^\prime + P_{\psi 2} \psi_2^\prime) \biggl) = 0. \ee For our third gauge we choose \be \label{A0gauge} \epsilon:=A_1 \approx 0, \ee which is the coulomb gauge with the constant, $A_1$ chosen to be zero. This condition, along with the electromagnetic constraint, \be \label{EMConstraint} e(P_{\psi 2} \psi_1 - P_{\psi 1} \psi_2)-P_{A1}^\prime \approx 0, \ee removes $A_1$ and its conjugate momentum $P_{A1}$ from the set of phase space variables. We can therefore set $\epsilon$ strongly to zero in the Hamiltonian as we did for the first two gauge choices. This gives the following Hamiltonian: \begin{align} H_{\rm total} =& \int dR \biggl[ N\biggl\{ -M^\prime +\frac{P_{\psi 1}^2 + P_{\psi 2}^2}{2{\cal A}_{n-2}R^{n-2}} +\frac{{\cal A}_{n-2}R^{n-2}}{2}(\psi_2^{\prime2} + \psi_1^{\prime2} ) \nonumber \\ & +\frac{P_{A1}^2}{2{\cal A}_{n-2}R^{n-2}} + \frac{N_r}{N} (P_{\psi 1} \psi_1^\prime + P_{\psi 2} \psi_2^\prime) \biggl\} + A_0\biggl(e(P_{\psi 2} \psi_1 - P_{\psi 1} \psi_2)-P_{A1}^\prime\biggl)\biggl], \label{HEMtotal2} \end{align} where it is understood that $P_{A1}$ is the solution of Eq.~(\ref{EMConstraint}). The consistency condition on Eq.~(\ref{A0gauge}) is given by \begin{align} \{\epsilon , H_{\rm total} \} \approx 0 &\to A_0^\prime + \frac{NP_{A1}}{{\cal A}_{n-2}R^{n-2}} \approx 0 \nonumber \\ &\to A_0^\prime \approx -\frac{eN}{{\cal A}_{n-2}R^{n-2}} \int dR (P_{\psi 2} \psi_1 - P_{\psi 1} \psi_2),\label{EM consistency condition 3} \end{align} which puts a condition on the final Lagrange multiplier and must be satisfied at every time slice. This is the last consistency condition on $\epsilon$. With the fully gauge fixed Hamiltonian (\ref{HEMtotal2}) we may write down Hamilton's equations of motion in terms of the remaining phase space variables, $\psi_1$, $P_{\psi1}$, $\psi_2$ and $P_{\psi2}$. The equations of motion are given by \begin{align} \label{EMpsi1EOM} \dot{\psi_1} =& N\left(\frac{P_{\psi1}}{{\cal A}_{n-2}R^{n-2}} + \frac{N_r}{N} \psi_1^\prime\right) - eA_0\psi_2,\\ \label{EMpsi2EOM} \dot{\psi_2} =& N\left(\frac{P_{\psi2}}{{\cal A}_{n-2}R^{n-2}} + \frac{N_r}{N} \psi_2^\prime\right) + eA_0\psi_1,\\ \label{EMP1EOM} \dot{P}_{\psi1} =& \left[N \left({\cal A}_{n-2}R^{n-2} \psi_1^\prime + \frac{N_r}{N} P_{\psi1} \right) \right]^\prime - eA_0 P_{\psi2},\\ \label{EMP2EOM} \dot{P}_{\psi2} =& \left[N \left({\cal A}_{n-2}R^{n-2} \psi_2^\prime +\frac{N_r}{N} P_{\psi2} \right) \right]^\prime + e A_0 P_{\psi1}. \end{align} It must be remembered that at every time slice the equations of motion must be supplemented by the consistency conditions (\ref{Nsig}), (\ref{EM consistency condition 2}), and (\ref{EM consistency condition 3}), as well as the Hamiltonian constraint: \be \label{EMNconstraint} -M^\prime + \frac{P_{\psi 1}^2 + P_{\psi 2}^2}{2{\cal A}_{n-2}R^{n-2}} + \frac{{\cal A}_{n-2}R^{n-2}}{2}(\psi_2^{\prime2} + \psi_1^{\prime2}) + \frac{P_{A1}^2}{2{\cal A}_{n-2}R^{n-2}} + \frac{N_r}{N} (P_{\psi 1} \psi_1^\prime + P_{\psi 2} \psi_2^\prime) = 0, \ee where it is understood that $P_{A1}$ is the solution of Eq.~(\ref{EMConstraint}). \section{Conclusions} \label{Conclusions} In this paper, we have performed the Hamiltonian analysis for spherically symmetric spacetimes in general Lovelock gravity in arbitrary dimensions. We have shown that, as in general relativity, the areal radius and the generalized Misner-Sharp quasi-local mass $M$ are natural canonical variables that yield the remarkably simple, geometrical action (\ref{eq:geometrodynamic L2}) for the generic theory. Using these variables also enabled us to rigorously derive the super-Hamiltonian and super-momentum constraints (\ref{eq:super H}) and (\ref{eq:super Hr}) for the most general theory, a task that would have been daunting at best, if not impossible, in terms of ADM variables. Most importantly, our results are useful: the geometrodynamic variables allow the physical phase space of the vacuum theory to be explicitly parametrized in terms of the ADM mass and its conjugate momentum, as done for general relativity by Kucha\v{r}~\cite{kuchar94}. This in turn provides a rigorous starting point for the quantization of Lovelock black holes using the techniques of \cite{Louko1996}. Finally, the simple form of the Hamiltonian allows us to gauge fix and derive the Hamiltonian equations of motion for the collapse of self gravitating matter in flat slice coordinates. We are now in a position to study the dynamics of black hole formation in generic Lovelock gravity. This is currently in progress. Finally we note that while the equations of motion remain virtually unchanged for the case of non-compact symmetry group (i.e. $k=0,-1$) we have nonetheless focused on spherical symmetry ($k=1$). This is perhaps the most interesting case because it describes physically relevant asymptotically flat black holes. The non-compact cases are also of interest in part because of the connection with the AdS/CFT correspondence, for example. However the boundary conditions for $k=0,-1$ require more detailed analysis. Moreover, it has only been rigorously proven that dimensional reduction commutes with the variational principle for spherically symmetric space-times. For these reasons we defer consideration of $k=0,-1$ to a separate work. {\bf Acknowledgments} HM thanks Jorge Zanelli for discussions. This research was supported in part by the Natural Sciences and Engineering Research Council of Canada and by the JSPS Grant-in-Aid for Scientific Research (A) (22244030). This work has been partially funded by the Fondecyt grants 1100328, 1100755 and by the Conicyt grant "Southern Theoretical Physics Laboratory" ACT-91. The Centro de Estudios Cient\'{\i}ficos (CECs) is funded by the Chilean Government through the Centers of Excellence Base Financing Program of Conicyt. TT thanks the University of Manitoba for funding. GK and TT are very grateful to CECs, Chile for its kind hospitality during the initial stages of this work. HM would like to thank University of Winnipeg, for its hospitality while part of this work was carried out. The authors thank Julio Oliva for bringing Ref.\cite{Palais1979} to their attention. GK is grateful to Jorma Louko for introducing him to the beauty of Kucha\v{r}' geometrodynamics. TT and GK also thank Danielle Leonard and Robert Mann for the collaboration in~\cite{TLKM} which was instrumental in leading to the present work.